In aquatic environments, hypoxia is a multi-dimensional stressor that can vary in O2 level (partial pressure of O2 in water, PwO2), rate of induction and duration. Natural hypoxic environments can therefore be very different from one another. For the many fish species that have evolved to cope with these different hypoxic environments, survival requires adjusting energy supply and demand pathways to maintain energy balance. The literature describes innumerable ways that fishes combine aerobic metabolism, anaerobic metabolism and metabolic rate depression (MRD) to accomplish this, but it is unknown whether the evolutionary paths leading to these different strategies are determined primarily by species' phylogenetic histories, genetic constraint or their native hypoxic environments. We explored this idea by devising a four-quadrant matrix that bins different aquatic hypoxic environments according to their duration and PwO2 characteristics. We then systematically mined the literature for well-studied species native to environments within each quadrant, and, for each of 10 case studies, described the species' total hypoxic response (THR), defined as its hypoxia-induced combination of sustained aerobic metabolism, enhanced anaerobic metabolism and MRD, encompassing also the mechanisms underlying these metabolic modes. Our analysis revealed that fishes use a wide range of THRs, but that distantly related species from environments within the same matrix quadrant have converged on similar THRs. For example, environments of moderately hypoxic PwO2 favoured predominantly aerobic THRs, whereas environments of severely hypoxic PwO2 favoured MRD. Capacity for aerial emergence as well as predation pressure (aquatic and aerial) also contributed to these responses, in addition to other biotic and abiotic factors. Generally, it appears that the particular type of hypoxia experienced by a fish plays a major role in shaping its particular THR.

Environmental hypoxia is a multi-dimensional stressor of many aquatic ecosystems, typically involving variations in O2 level (partial pressure of O2 in water, PwO2), rate of hypoxia induction and duration of exposure. Aquatic hypoxic environments can therefore differ from one another considerably. For example, intertidal pools oscillate between ∼80 and 0 kPa PwO2 each day (Richards, 2011), whereas oceanic oxygen minimum zones (OMZs), which typically occur at depths of 200 to 1500 m, remain at stable hypoxic PwO2 (≤4.2 kPa; Seibel, 2011). Orders of fishes have independently evolved abilities to tolerate and exploit different hypoxic environments, and the literature is rife with studies that reveal the different strategies that allow them to do this. What remains unknown, however, is whether similar hypoxic environments beget similar hypoxic survival strategies, independent of phylogenetic relationships.

Our Review will explore variation in the hypoxic survival strategies of different fish species in the context of these species' native hypoxic environments. We will represent each species' hypoxic survival strategy using the total hypoxic response (THR; see Glossary), which we define as the combination of these three metabolic modes an animal uses in hypoxia: (i) sustained aerobic metabolism via a wide array of mechanisms that enhance capacities for O2 uptake, transport and delivery; (ii) increased use of anaerobic metabolism; and (iii) metabolic rate depression (MRD; see Glossary). The THR is useful in this respect because it is a complex phenotype that is contributed to by numerous traits operating at lower levels of organization. Although these traits can vary among species [e.g. some species increase haemoglobin (Hb)–O2 binding affinity, whereas others increase haematocrit (see Glossary)] or be similar but achieved through different mechanisms (e.g. high gill surface area that is constitutively expressed or increased via remodeling/plasticity), they may result in consistent responses at higher levels of organization (e.g. sustained aerobic metabolism). Interpreting hypoxic responses at these higher levels of biological organization may reveal patterns of convergent evolution (see Glossary) that are less visible at lower levels.

List of symbols and abbreviations
     
  • ASR

     aquatic surface respiration

  •  
  • CrP

     creatine phosphate

  •  
  • Hb

     haemoglobin

  •  
  • Hb P50

     partial pressure of O2 at which Hb is 50% saturated with O2

  •  
  • O2

     oxygen consumption rate

  •  
  • MRD

     metabolic rate depression

  •  
  • OMZ

     oxygen minimum zone

  •  
  • PwO2 

     partial pressure of oxygen in water

  •  
  • THR

     total hypoxic response

Glossary

Aquatic surface respiration

A breathing technique that involves skimming the relatively well-oxygenated surface layer of the water column.

Bohr effect

A reduced haemoglobin–oxygen binding affinity resulting from a reduced pH.

Convergent evolution

Independent evolution of similar traits in different evolutionary lineages.

Eutrophic

A body of water enriched with nutrients and minerals, resulting in excessive growth of plants and algae.

Haematocrit

Proportion of whole blood that consists of red blood cells.

Macrophyte

An aquatic plant that grows in or near water, and is emergent, submergent or floating.

Metabolic acidosis

Reduced pH resulting from metabolic activity.

Metabolic rate depression

A coordinated and reversible depression of metabolic rate below standard metabolic rate.

Oligotrophic

A body of water with low accumulation of nutrients and minerals, resulting in sparse growth of plants an algae.

Critical oxygen tension (Pcrit)

The lowest PO2 at which an animal can sustain routine O2, below which O2 progressively decreases.

Pleiotropy

Phenotypic effect of a gene on more than one trait.

Total hypoxic response

The hypoxia-induced combination of sustained aerobic metabolism, enhanced anaerobic metabolism and metabolic rate depression (encompassing also the mechanisms underlying these metabolic modes) that an animal uses to survive hypoxia.

Ventilation amplitude

Volume of water pumped over the gills during a single breath.

Ventilation rate

Number of breaths per minute.

Water column stratification

Vertical distribution of water layers that form as a result of low mixing between water masses of different properties (e.g. hypoxia, salinity, density, temperature).

White muscle

A rapid or fast-twitch muscle fibre type.

There are costs associated with hypoxic survival strategies

Although the metabolic strategies that contribute to the THR – aerobic metabolism, anaerobic metabolism and MRD – benefit hypoxic survival (Fig. 1; and see Box 1 for descriptions of these metabolic modes and the data that may be used to interpret them), each strategy has limits and potential costs that may influence how it is (or is not) used in certain hypoxic environments. The main limit of aerobic metabolism is a critical PwO2 (Pcrit; see Glossary) below which aerobic reliance becomes severely compromised (see Box 1, Fig. 1). Capacity for aerobic metabolism may be improved through different mechanisms of increasing O2 supply, but these too can come at a cost. For example, increasing respiratory surface area aids in oxygen extraction from the water, but with potential negative consequences of compromising ion and water fluxes, increasing detrimental uptake of ammonia and other toxic substances, and increasing the likelihood of bleeding (reviewed in Nilsson and Randall, 2010). Similarly, an increase in ventilation and haematocrit can improve O2 extraction and delivery, but an increase in ventilation incurs a significant energetic cost (Hughes and Sunders, 1970; Steffensen, 1985), and high haematocrit increases blood viscosity, thus increasing blood flow resistance (reviewed in Gallaugher and Farrell, 1998).

Fig. 1.

The aerobic, anaerobic and metabolic rate depression (MRD) responses of a fish to hypoxia. (A) The aerobic response of a typical fish to decreasing partial pressure of oxygen in water (PwO2). The solid curve represents the oxygen consumption rate (O2) that is required to support routine metabolic rate, and a typical fish can maintain this O2 over a broad PwO2 range (including moderately hypoxic PwO2) using the mechanisms listed on the right. The ability to use these mechanisms varies among species. Pcrit (vertical dashed line) is the critical PwO2 at which the animal is unable to extract sufficient environmental O2 to support routine O2. PwO2 values below Pcrit are severely hypoxic, and, because aerobic ATP supply is compromised at these PwO2 values, the fish's survival depends on enhanced reliance on anaerobic ATP supply pathways (glycolysis) and/or MRD. (B) Theoretical metabolic responses of an animal at sub-Pcrit (i.e. severely hypoxic) PwO2 as a function of time, with the y-axis representing the rate of energy use instead of O2. Line 1 (solid line) represents a fish that maintains routine metabolic rate in severe hypoxia by supplementing its reduced aerobic ATP supply capacity with an enhanced reliance on glycolysis. Line 2 (dashed line) represents a fish that induces MRD to better match its ATP demand to its reduced ATP supply capacity, and this is accomplished by reducing processes such as those listed on the right. Both metabolic responses to severe hypoxia are ultimately unsustainable, but the MRD response (line 2), here illustrated with its approximately 50% depression of metabolic rate, allows this hypothetical fish to survive for twice as long in severe hypoxia.

Fig. 1.

The aerobic, anaerobic and metabolic rate depression (MRD) responses of a fish to hypoxia. (A) The aerobic response of a typical fish to decreasing partial pressure of oxygen in water (PwO2). The solid curve represents the oxygen consumption rate (O2) that is required to support routine metabolic rate, and a typical fish can maintain this O2 over a broad PwO2 range (including moderately hypoxic PwO2) using the mechanisms listed on the right. The ability to use these mechanisms varies among species. Pcrit (vertical dashed line) is the critical PwO2 at which the animal is unable to extract sufficient environmental O2 to support routine O2. PwO2 values below Pcrit are severely hypoxic, and, because aerobic ATP supply is compromised at these PwO2 values, the fish's survival depends on enhanced reliance on anaerobic ATP supply pathways (glycolysis) and/or MRD. (B) Theoretical metabolic responses of an animal at sub-Pcrit (i.e. severely hypoxic) PwO2 as a function of time, with the y-axis representing the rate of energy use instead of O2. Line 1 (solid line) represents a fish that maintains routine metabolic rate in severe hypoxia by supplementing its reduced aerobic ATP supply capacity with an enhanced reliance on glycolysis. Line 2 (dashed line) represents a fish that induces MRD to better match its ATP demand to its reduced ATP supply capacity, and this is accomplished by reducing processes such as those listed on the right. Both metabolic responses to severe hypoxia are ultimately unsustainable, but the MRD response (line 2), here illustrated with its approximately 50% depression of metabolic rate, allows this hypothetical fish to survive for twice as long in severe hypoxia.

Box 1. The metabolic responses of fishes to hypoxia

Energy metabolism

Energy metabolism is the rate of ATP turnover of a cell, tissue or whole animal (or organism). ATP is supplied by oxidative phosphorylation and/or anaerobic glycolysis, and is consumed by energy-demanding biological processes ranging from whole animal behaviour to protein translation. For a hypoxia-exposed animal, energy metabolism is the combined sum of aerobic metabolism, anaerobic metabolism and metabolic rate depression.

Aerobic metabolism

Aerobic metabolism centers on oxidative phosphorylation, the O2-dependent process by which ATP is produced in the mitochondria. Optimizing steps along the O2-transport cascade to more efficiently move O2 from the environment to the mitochondria can aid hypoxic survival (Weibel, 1984), and hypoxia-adapted fishes have evolved traits at each step to do so (Sollid et al., 2003; Gracey et al., 2001; Affonso et al., 2002; Lai et al., 2006; Turko et al., 2014; Holeton and Randall, 1967a,b; Itazawa and Takeda, 1978; Tzaneva et al., 2011; Vulesevic and Perry, 2006; Sundin et al., 1995). The collective effectiveness of these traits to extract O2 from hypoxic water is quantified by the critical PwO2 (Pcrit; Fig. 1) of O2 uptake rate (O2). The lower the Pcrit, the greater the PwO2 range over which the fish can maintain routine metabolic rate aerobically. As PwO2 drops below Pcrit and aerobic metabolism becomes compromised, a fish's survival requires either supplementing aerobic metabolism with anaerobic metabolism to maintain routine metabolic rate, or depressing ATP-consuming processes to better match reduced ATP supply rates.

Data that are relevant to aerobic metabolism

O2; Pcrit ; behaviours that maximize O2 uptake (from water or air); ventilatory responses; respiratory surface anatomy; haematology and Hb function; circulatory anatomy and physiology; aerobic enzyme function; mitochondrial function.

Anaerobic metabolism

Anaerobic metabolism includes anaerobic glycolysis and creatine phosphate (CrP) hydrolysis. CrP is important for activity, but because CrP reserves are small and quickly depleted, they play a minimal role in hypoxic survival. Glycolysis is beneficial in O2-limited environments because it allows for an O2-independent supply of ATP. Most fishes, tolerant and intolerant, upregulate glycolysis during hypoxia, but tolerant fishes display a suite of biochemical adaptations that collectively enhance their anaerobic potentials (Hochachka and Somero, 2002; Farwell et al., 2007; Mandic et al., 2013; Lushchak et al., 1998; Abbaraju and Rees, 2011; Johnston, 1977; Shoubridge and Hochachka, 1980). However, even in hypoxia-tolerant species, should the hypoxic exposure last too long, glycolysis cannot sustain sufficient ATP supply and survival becomes dependent on the reduction of metabolic demand.

Data that are relevant to anaerobic metabolism

A species' reliance on anaerobic glycolysis may be represented by tissue/plasma lactate accumulation, tissue/whole-body glycogen depletion, tissue/plasma ethanol accumulation, or ethanol excretion to the water; tissue/whole-body [glycogen] quantifies a species' anaerobic fuel stores; glycolytic enzyme function represents potential glycolytic flux rates.

Metabolic rate depression

Metabolic rate depression (MRD) is a reduction in metabolic rate below standard metabolic rate (SMR; Hochachka and Somero, 2002; Richards, 2009). Metabolic rate is reduced to SMR through adjustments at behavioural and physiological levels, and below SMR (i.e. MRD) through adjustments at physiological and cellular/biochemical levels (Brett and Groves, 1979; Chiba, 1983; Pedersen, 1987; Nilsson et al., 1993; Schurmann and Steffensen, 1994; McKenzie et al., 1995; Hochachka et al., 1996; Hochachka and Somero, 2002; Thomas et al., 2006; Fitzgibbon et al., 2007; Perry et al., 2009; Richards, 2009; Wang et al., 2009; Wu, 2009). MRD is an effective mechanism for maintaining energy balance when ATP supply is limited at sub-PcritPwO2, but although some fishes use it, others do not. Those that do use MRD tend to be highly tolerant.

Data that are relevant to MRD

Calorimetric measurements of metabolic heat production, though rare in the fish literature, are the best indicator of MRD because they inherently account for aerobic and anaerobic contributions to metabolic rate (Nelson, 2016); simultaneous measurements of O2 and anaerobic reliance (see above); reduced rates of ATP-consuming processes such as ATPase activity and protein translation may (but not necessarily) indicate MRD; whole-body quiescence reduces ATP demand, though is not MRD because it does not reduce metabolic rate below standard levels.

For anaerobic metabolism (i.e. anaerobic glycolysis, hereafter referred to as glycolysis), the main limit is a finite substrate pool (glucose, glycogen) that restricts the time over which an animal may rely on glycolysis (Wang et al., 2009). The main cost of prolonged anaerobic reliance is a metabolic acidosis (see Glossary) that jeopardizes the fish's health and hypoxia tolerance (Driedzic and Gesser, 1994; Hochachka and Somero, 2002; Nilsson and Östlund-Nilsson, 2008). And more generally, glycolysis is relatively inefficient at converting energy stored in food to forms that are usable by the cell (ATP).

For MRD, the costs can be physiological or ecological. Physiologically, these include oxidative damage (Carey et al., 2000), reduced growth, repair and immunocompetence (Burton and Reichman, 1999), and cognitive impairments stemming from neuronal damage (Popov et al., 1992; Lefevre et al., 2017). Ecologically, costs include ceased reproduction (Humphries et al., 2003) and increased susceptibility to predation stemming from significantly reduced awareness and motor activity (Draud et al., 2004; Kokurewicz, 2004; Lanszki et al., 2006; Wiklund et al., 2008; Estok et al., 2009; Sommer et al., 2009; Olofsson et al., 2011).

There is great variation in how different fish species use each metabolic strategy in response to hypoxia. Some species prioritize aerobic and anaerobic metabolism (e.g. intertidal sculpins), whereas others rely on anaerobic metabolism coupled with MRD (e.g. crucian carp). It is possible that the diversity of THRs is partly attributable to the aforementioned costs, which may be more or less relevant to a given species depending on its ability to mitigate them. This ability may be influenced by factors such as genetic constraint or the ecological environment. However, no study has explicitly examined the causes underlying why different hypoxia-tolerant species have (or have not) evolved particular THR profiles.

THRs may arise through different mechanisms

Despite multiple THRs arising among the hypoxia-tolerant fishes, there is a finite number of evolutionary solutions and most species can be broadly grouped into clusters of similar metabolic strategies. Three mechanisms may influence this probability of convergence: genetic constraints, phylogenetic history and/or natural selection (Christin et al., 2010; Losos, 2011; Rosenblum et al., 2014).

Genetic constraints, such as pleiotropy (see Glossary) or limited genetic variation, increase the likelihood of repeated evolution of a given phenotype because the constraints decrease the number of evolutionary paths available in response to an environmental stressor such as hypoxia (Chevin et al., 2010; Conte et al., 2012). Phylogenetic history may also lead to similar THRs, whereby the shared genetic backgrounds of closely related species increase their probabilities of evolving parallel solutions (Rosenblum et al., 2014). Natural selection may also be a possible driver of convergence, where the ecological environment is the primary determinant of evolution of a similar THR across species. If the environment is a driving factor in the evolution of a particular THR, then mining the extensive hypoxia literature should reveal a pattern whereby distantly related species native to similar hypoxic environments display similar THRs. Conversely, closely related species native to different hypoxic environments may display different THRs.

Exploring the role of the environment in THR evolution

To begin to explore how variation in the hypoxic environment influences the THR of fishes, we first devised a four-quadrant matrix (Q1–Q4; Fig. 2) that bins different aquatic hypoxic environments according to their time and PwO2 dimensions: Q1, hypoxia that is short in duration and moderate in PwO2; Q2, short in duration and severe in PwO2; Q3, long in duration and moderate in PwO2; and Q4, long in duration and severe in PwO2. Although this broad binning has its shortcomings (e.g. it is insensitive to finer-scale environment variations), it is to our knowledge the first meta-analysis to incorporate this level of hypoxic complexity, a necessary step towards understanding the fundamental aspects of hypoxia tolerance.

We then turned to the literature for examples of particularly well-studied species native to environments within these four quadrants. We found 10 hypoxic environment types with well-studied resident species (or multi-species systems) with respect to THR-related characteristics. Other relevant species–environment combinations exist in both the literature and the natural world, but we chose ours based on well-resolved understandings of the O2 dynamics of the environment and an understanding of at least two of the three metabolic modes of the THR. We mined the literature systematically (see Table S1), but were limited in some cases by information availability (especially regarding MRD, the least studied metabolic mode) and the possibility of ascertainment bias regarding some traits and species (i.e. certain phenomena are better suited to study in some species than others). This may have influenced some of our interpretations. However, for the most part, compiling results from across multiple studies allowed us to piece together most species' THRs, something that had not been done before. Interpreting these THRs in the context of each species' natural hypoxic environment revealed patterns whereby distantly related species under similar hypoxic pressures rely on similar THRs, and suggested that other factors such as predation also likely play roles in shaping the THR.

The following sections are organized by quadrant, and within each section are descriptions of that quadrant's representative case studies. Although we have elected to describe Q1 below, species belonging to these environments are not particularly hypoxia-tolerant. As such, they are not the focus of the Review but have been included as a comparison case of how mildly tolerant species may respond to hypoxia. We have detailed our literature-mining process in Table S1, and have also included an extensive outline of these studies and their reported THR-related values. In the final section of the paper, we summarize our findings and reflect on what they might mean for the evolution of hypoxic survival strategies.

Many aquatic species encounter low PwO2 in the environment, but most avoid or escape these hypoxic zones and are therefore only transiently exposed to moderate hypoxia. One such species is the well-studied rainbow trout (Oncorhynchus mykiss), an inhabitant of streams and lakes, where pockets of O2-depleted water periodically form. Like most salmonids, rainbow trout are not particularly hypoxia tolerant (Doudoroff and Shumway, 1970) and avoid hypoxic areas by migrating vertically away from the reduced PwO2 of deeper waters (Rowe and Chisnall, 1995). Studies performed under controlled laboratory conditions have found that, similar to hypoxia-tolerant species (discussed in subsequent sections), rainbow trout rapidly (minutes to hours) attempt to maximize O2 uptake during hypoxia by increasing ventilation rate (Holeton and Randall, 1967a,b) and amplitude (Hughes and Sunders, 1970) (see Glossary), and blood Hb–O2 affinity via reduced intraerythrocytic ATP concentrations (Tetens and Lykkeboe, 1981). In contrast to tolerant species that typically maintain O2 consumption rate (O2) as PwO2 drops towards Pcrit, rainbow trout increase O2, likely the result of enhanced ventilatory efforts (Hughes and Saunders, 1970). Further PwO2 reductions see the trout enhance anaerobic metabolism (Dunn and Hochachka, 1986), but, unlike most hypoxia-tolerant fishes, the trout's anaerobic metabolism does not appear to be fuelled by hepatic glycogen stores (Dunn and Hochachka, 1987; Van Raaij et al., 1996). And perhaps most indicative of the trout's relative intolerance, hypoxia exposure rapidly and significantly reduces cellular [ATP], [creatine phosphate] ([CrP]) and energy charge in crucial tissues such as the heart (Dunn and Hochachka, 1986).

The rainbow trout therefore employs a primarily aerobic THR with some contribution of anaerobic metabolism in severe hypoxia. This is similar to many species (see below), but the trout's relatively limited capacity for these strategies results in cellular [ATP] imbalance and, consequently, a low hypoxia tolerance. Nevertheless, given the moderate PwO2 and escapable nature of the trout's hypoxic environment, this THR is sufficient.

Intertidal zones

Fishes routinely found in high intertidal pools experience dramatic and acute changes in PwO2, unlike their subtidal counterparts that inhabit much more O2-stable environments (see environment description in Box 2). Intertidal sculpins, such as the tidepool sculpin (Oligocottus maculosus) and the intertidal triplefin twister (Bellapiscis medius), show lower Pcrit values than do subtidal sculpins and triplefins, respectively (Hilton et al., 2008; Mandic et al., 2009a). The intertidal fishes' lower Pcrit, which indicates their ability to maintain routine O2 to lower PwO2, is achieved through behavioural and physiological modifications to the O2 cascade that maximize O2 extraction and delivery to the tissues, reducing the impact of environmental O2 limitation on aerobic respiration.

Box 2.Aquatic hypoxic environments

Each aquatic hypoxic environment is unique in the way physical and biological factors create its hypoxic events. However, we have binned 12 different hypoxic environments according to their PwO2 and durations over which they remain hypoxic using a four-quadrant matrix: Q1, hypoxia that is short in duration and moderate in PwO2; Q2, short duration and severe; Q3, long duration and moderate; Q4, long duration and severe. We describe these environments below.

Q1: Thermally stratified lakes

The hypolimnion is the lower layer of a thermally stratified lake, and is frequently hypoxic, especially during the summer months (Rowe and Chisnall, 1995; Roberts et al., 2009).

Q2: Intertidal zones

Tidepools high in the intertidal zone become isolated from the ocean for hours to days with the falling tide. Their small water volumes and often-dense biota result in enormous fluctuations in PwO2 reaching anoxia at night and up to 80 kPa in the day (Truchot and Duhamel-Jouve, 1980; Burggren and Roberts, 1991; Richards, 2011).

Q2: Estuaries

Estuarine waters are subject to tides and strong winds that upwell O2-poor bottom waters and cause severe diel variation in PwO2 levels, with hypoxic events often occurring at night (Breitburg, 1990; D'Avanzo and Kremer, 1994). Dense biota and water column stratification (see Glossary) further reduce PwO2 (Breitburg, 1990).

Q2: Coral reefs

Pools on coral reef flats become hypoxic when isolated during tidal cycles, and nocturnal O2 consumption among the coral colonies themselves can severely reduce PwO2 to as low as 0.4 kPa (Goldshmid et al., 2004; Nilsson and Ostlund-Nilsson, 2006; Nilsson et al., 2007).

Q2, Q3: Oceanic oxygen minimum zones (OMZs)

OMZs occur at depths between 200 and 1500 m, where certain biological and physical processes combine to reduce PwO2 to <6.4 kPa around the OMZ's periphery and 0.5 kPa in its centre. Biologically, a high density of aerobic bacteria reduce the OMZ's O2 levels as they feed upon the organic matter falling from the mixed layer above, while physically, a lack of atmospheric contact and low levels of convective mixing keep these waters low in O2.

Q3: High-altitude lakes

Atmospheric pressure decreases by ∼0.91 kPa with every 100 m of altitudinal ascent. Water bodies at higher altitudes consequently have lower partial pressures for all gases, including O2. These hypoxic exposures are typically chronic and moderate in PwO2, reaching ∼14 kPa as a result of altitude in lakes at 3750 m above sea level (about the highest at which fish species have been studied).

Q3, Q4: Winterfreeze lakes

Winterfreeze lakes generally occur at high elevations or at far northern or southern latitudes, where low wintertime atmospheric temperatures freeze the lake's surface layer. Biological activity reduces PwO2, and the ice/snow layer prevents photosynthesis and water–atmosphere mixing until the spring thaw (Ultsch, 1989; Barica and Mathias, 1979; Mathias and Barica, 1980). The relatively low biological activity levels of oligotrophic lakes (Fig. 2, Q3) can cause moderately hypoxic PwO2, whereas the high activity levels of eutrophic lakes can cause anoxia for months (Fig. 2, Q4) (Vornanen, 2004).

Q4: Swamps

Swamps, such as those surrounding Lake Victoria, typically experience hypoxia that is chronically low in PwO2, varying from 0.6 kPa in the bottom layers at night to a maximum of 3.2 kPa in the upper layer during the day (Chapman et al., 2002) as a result of high biological activity.

Q2, Q3, Q4: The Amazon

The Amazon basin floods each year when the Amazon River spills over its riverbanks into the surrounding forests (igapo) and floodplains (varzea), bringing many of the Amazon's 5600+ fish species with it (Albert and Reis, 2011). The basin is a network of complex heterogeneous environments where interacting biological and physical factors create long- and short-term fluctuations in O2 levels (Val et al., 1998). At the peak of the wet season, all of the flooded areas are interconnected, allowing fish to move among them. The hypoxia that occurs during this season arises from extensive floating macrophytes (see Glossary) and occurs mainly in the varzea lakes (Val and Almeida-Val, 1995). As the season wears on, the water levels recede and leave behind smaller, isolated water bodies that become hypoxic (even anoxic) as a result of high biological activity and a lack of light penetration (Val and Almeida-Val, 1995; Val et al., 1998). These habitats can remain deeply hypoxic for months at a time, even chronically, or undergo large diurnal changes in O2, all depending on the shape, size, depth, winds and vegetation of the varzea lakes (Val and Almeida-Val, 1995).

Fig. 2.

An environmental hypoxia matrix. Natural hypoxic environments are grouped according to their O2 levels (PwO2) and time durations over which they remain hypoxic. Quadrant 1 (Q1) portrays moderately hypoxic PwO2+short-duration environments. Q2 portrays severely hypoxic PwO2 + short-duration environments. Q3 portrays moderately hypoxic PwO2 + long-duration environments. Q4 portrays severely hypoxic PwO2 + long-duration environments. Note that as the Amazon basin encompasses a wide diversity of hypoxic habitats, it is listed under Q2–Q4 (for greater description, see ‘Heterogeneous hypoxic environments’ section).

Fig. 2.

An environmental hypoxia matrix. Natural hypoxic environments are grouped according to their O2 levels (PwO2) and time durations over which they remain hypoxic. Quadrant 1 (Q1) portrays moderately hypoxic PwO2+short-duration environments. Q2 portrays severely hypoxic PwO2 + short-duration environments. Q3 portrays moderately hypoxic PwO2 + long-duration environments. Q4 portrays severely hypoxic PwO2 + long-duration environments. Note that as the Amazon basin encompasses a wide diversity of hypoxic habitats, it is listed under Q2–Q4 (for greater description, see ‘Heterogeneous hypoxic environments’ section).

Behaviourally, intertidal fishes respond to hypoxia with aquatic surface respiration (ASR; skimming of the relatively well-oxygenated surface layer of the water column; see Glossary) and aerial emergence, where the fish leave the tidepool to respire in air (reviewed in Bridges, 1988; Martin, 1995). Initiation of ASR and/or emergence occurs as PwO2 falls below Pcrit (Congleton, 1980; Innes and Wells, 1985; Hill et al., 1996; Mandic et al., 2009b). If intertidal fishes are restricted from the surface water or air, O2 continually declines with a decrease in PwO2 below Pcrit; however, given the opportunity to access surface water or air, intertidal fish will use ASR and/or emerge to breathe air, thus maintaining routine O2 (Yoshiyama and Cech, 1994; Martin, 1996). Indeed, a number of studies report similar or only slightly reduced respiratory rates in air compared with in water (Wright and Raymond, 1978; Daxboeck and Heming, 1982; Martin, 1991; Yoshiyama and Cech, 1994; Sloman et al., 2008). These behavioural responses uncouple the intertidal fishes from their aquatic habitat and allow them to maintain O2 by accessing the well-oxygenated upper layer of water or air when the bulk water of a tidepool becomes severely hypoxic (i.e. sub-Pcrit) (Yoshiyama and Cech, 1994; Martin, 1996).

Although these behaviours enhance hypoxic survival, they also significantly elevate the fishes' risk of aerial predation (Kramer, 1983). A perceived predatory threat from above will send the intertidal fishes back into the tidepool's severely hypoxic water (or delay their emergence from it; Hugie et al., 1991; Shingles et al., 2005; Sloman et al., 2008), and in these scenarios, survival ultimately depends on a suite of physiological and biochemical adaptations (Brix et al., 1999; Mandic et al., 2009a; Craig et al., 2014; Lau et al., 2017). Compared with their subtidal counterparts, intertidal fishes exhibit: higher mass-specific gill surface area (Mandic et al., 2009a); higher haematocrits and blood–O2 carrying capacities (Craig et al., 2014); higher Hb–O2 affinities (Mandic et al., 2009a; Brix et al., 1999, respectively); and higher cytochrome c oxidase O2 affinities (Lau et al., 2017). These adaptations allow intertidal fishes to rely primarily on aerobic metabolism even if predatory threats deny them surface access for periods of time. However, if PwO2 in the tidepool falls below Pcrit, then the fishes' survival may depend on anaerobic metabolism and/or MRD.

Intertidal sculpins have high capacities for anaerobic glycolysis. Specifically, their glycogen reserves are large and their glycolytic enzyme activity levels in brain are significantly higher than those of closely related subtidal sculpins (Mandic et al., 2013). Similarly, plainfin midshipman males, which are exposed to repeated hypoxic bouts while tending their nests, have higher glycogen reserves (liver) and glycolytic enzyme capacities (gill, skeletal muscle) than the less-tolerant females, which do not tend the nests (LeMoine et al., 2014). And both tidepool sculpins and the plainfin midshipman exhibit significant accumulation of plasma lactate when exposed to ecologically relevant hypoxic bouts lasting 4 to 6 h, indicating the activation of anaerobic glycolysis (Speers-Roesch et al., 2013; Craig et al., 2014).

Scant information exists on MRD in intertidal fishes. It is known that tidepool and rosylip sculpins (Ascelichthys rhodorus), both intertidal species, reduce locomotor activity and enter quiescent states when denied access to air, a measure that helps conserve energy (Yoshiyama et al., 1995). However, it is not known whether cellular MRD occurs during this quiescent state. Parental male plainfin midshipman have been suggested to induce cellular MRD during hypoxia (Craig et al., 2014), but the evidence is based solely on reduced Na+/K+-ATPase activity levels in the gill (and no decrease in liver). It may be that the intertidal environment, which is rich in predators, may not favour MRD owing to the reduced predator avoidance abilities that accompany a metabolically depressed state. In any case, careful work on these intertidal species' MRD use would compliment the extensive work that has been done on their aerobic and anaerobic hypoxic defenses.

Taken together, the available evidence thus far suggests that intertidal fishes prioritize aerobic metabolism under all possible hypoxic conditions, and likely rely on anaerobic glycolysis (and perhaps not MRD) when forced to spend time in severely hypoxic water.

Estuaries

Estuarine fishes experience O2 regimes (severe PwO2 that is short in duration; see Box 2) and aerial predation pressures similar to those of intertidal fishes (Kneib, 1982; Burnett et al., 2007). Two well-studied estuarine species are the Atlantic killifish (Fundulus heteroclitus) and the gulf killifish (Fundulus grandis), whose Pcrit values are similar to those of the intertidal fishes (see Table S1; Cochran and Burnett, 1996; Virani and Rees, 2000; McBryan et al., 2016). As estuarine PwO2 falls, both Fundulus species typically skim the water's surface and perform ASR (Wannamaker and Rice, 2000; Love and Rees, 2002), an important behavioural mechanism that contributes to alleviating the negative effect of hypoxia on growth rate in gulf killifish (Stierhoff et al., 2003). Akin to the intertidal fishes, the killifishes' THR is predominantly aerobic and anaerobic metabolism rather than MRD, although how these metabolic strategies are combined depends on the hypoxic time frame.

During initial hypoxia exposures, there is an increase in transcription of genes associated with oxidative phosphorylation, suggesting that enhanced aerobic enzyme activity is among the first lines of metabolic defense in the gulf killifish (Everett et al., 2012). In both killifish species, sub-PcritPwO2 levels activate anaerobic metabolism in the liver and the white muscle (see Glossary) during short-term hypoxia (Virani and Rees, 2000; Richards et al., 2008), but prolonged exposure (days to weeks) causes a shift to reliance on MRD in the white muscle but not in the liver (Kraemer and Schulte, 2004; Martinez et al., 2006; Richards et al., 2008; Abbaraju and Rees, 2011). This indicates that the type of hypoxia strongly influences the metabolic response of estuarine fishes. Borowiec et al. (2015) explored this idea by acclimating Atlantic killifish to either chronic or intermittent (diel cycles) hypoxia exposures. Both exposure types lowered Pcrit and routine O2, but only killifish acclimated to intermittent hypoxia upregulated mechanisms that enhance glycolytic capacity and the processing of glycolytic end-products. Killifish acclimated to chronic hypoxia exhibited modified gill morphology (reduced filament length; reduced mitochondrion-rich cells that potentially decrease ion loss and cost of osmoregulation) in a way that may decrease metabolic demand (Borowiec et al., 2015).

Overall, it appears that killifish exposed to short-term intermittent hypoxia rely on an aerobic-anaerobic THR, while those exposed to long-term hypoxia rely less on anaerobic metabolism and perhaps more on MRD. Given that the killifishes' natural habitat typically experiences tidally influenced intermittent hypoxia, the former is likely the predominant THR, similar to the intertidal fishes. This THR may be well suited – and perhaps selected for – in predator-rich environments that experience rapid and severe fluctuations in PwO2.

Coral reefs

Hypoxic events in coral reefs tend to be severe, cyclical and short in duration (see Box 2), and so we would predict THRs among coral reef inhabitants that are similar to those of intertidal and estuarine fishes. Generally, coral reef fishes have relatively low Pcrit (3.1–6.1 kPa; Nilsson and Ostlund-Nilsson, 2004; Nilsson et al., 2004; Wong et al., 2018), but in-depth THR information is scant. The best-studied species in this regard is the epaulette shark (Hemischyllium ocellatum), a reef flat inhabitant. The shark's Pcrit is approximately 5.1 kPa (Speers-Roesch et al., 2012a), the lowest of any elasmobranch tested (Routley et al., 2002; Speers-Roesch et al., 2012a) and similar to those of teleost reef inhabitants. A comparison with a much less-tolerant elasmobranch, the shovelnose ray (Aptychotrema rostrata), revealed the epaulette shark to have a higher Hb–O2 affinity, higher arterial blood O2 content, higher O2 at sub-PcritPwO2 and better maintained routine cardiovascular function in hypoxia (Speers-Roesch et al., 2012a,b; Hickey et al., 2012), suggesting an aerobic-focused THR. Furthermore, elevated ventilatory frequencies (Routley et al., 2002) and altered blood flow patterns that enhance blood supply to the gills and return it directly to the heart (Stenslokken et al., 2004) support the epaulette shark's aerobic-focused THR.

Though epaulette sharks are known to clamber over land in search of water when the receding tide exposes tidal flats (Goto et al., 1999), they are not known to exploit the air's (or surface waters') higher O2 levels like intertidal and estuarine fishes. When PwO2 falls to ∼3.7 kPa (i.e. sub-Pcrit) as it typically does each night (Routley et al., 2002), anaerobic metabolism contributes to ATP production, as indicated by significant lactate accumulation with progressive, cyclical hypoxia exposures (Wise et al., 1998; Routley et al., 2002). There is conflicting information about whether the shark also uses MRD in these situations. Evidence for MRD use includes the shark's loss of righting reflex with anoxia exposure despite stable brain ATP levels (Renshaw et al., 2002), and reduced neuronal oxidative demand with cyclical severe hypoxia exposure (Mulvey and Renshaw, 2000). Relatively low cardiac lactate levels of hypoxia-exposed epaulette sharks are speculated to be due to MRD in extra-cardiac tissues, which would leave more O2 available for the heart (Speers-Roesch et al., 2012b). However, other studies have found no evidence of MRD in epaulette sharks (Dowd et al., 2010), nor a loss of body posture or voluntary movement with cyclical severe hypoxia (Wise et al., 1998). The conflicting results are possibly a function of the studies' different experimental hypoxia exposure protocols, but nevertheless, the presence of MRD in even some of these studies suggests that the epaulette shark has evolved an ability to use MRD, and perhaps does so in the wild. This is unlike the previously discussed intertidal and estuarine species, which are not known to employ MRD when exposed to tidally relevant (i.e. short-term intermittent) hypoxia despite their similar natural hypoxic habitats (though killifish may do so when acclimated to chronic hypoxia; see above). We speculate that the reason may involve predation pressure. Whereas the previous species are small and subject to predation during their hypoxia exposures, the shark is relatively large and less likely to encounter predators during its hypoxia exposures. Thus, the reduced predation risk may allow the shark to depress metabolism during hypoxia.

Oceanic OMZs – migratory residents

There are two types of OMZ resident: permanent and migratory. Permanent residents, which we discuss later, spend their entire lives in the OMZ and therefore experience chronic moderate-to-severe PwO2 levels (Q3 and Q4 in Fig. 2). Migratory residents, by contrast, spend their days in the OMZ's severely hypoxic centre and migrate vertically into well-oxygenated surface waters each night to feed in the cover of darkness (Seibel, 2011). This migratory pattern exposes these animals to progressively changing PwO2, becoming normoxic with upwards migration and hypoxic with downwards migration. The hypoxic exposures experienced by these animals are therefore severe (PwO2) and short in duration (see Box 2), similar to those described above. Despite this similarity, migratory OMZ residents use a different THR. Although they tend to possess traits that enhance O2 extraction (e.g. Seibel, 2013; Trueblood and Seibel, 2013) and glycolytic capacity (e.g. Gonzalez and Quiñones, 2002; Torres et al., 2012), migratory OMZ residents rely primarily on MRD while in the deeply hypoxic OMZ during the day (Seibel, 2011; Seibel et al., 2016). For example, the jumbo (or Humboldt) squid (Dosidicus gigas) depresses metabolic rate by 87% when held at 0.6 kPa, the PwO2 at which it typically spends the daytime in the OMZ (Rosa and Seibel, 2010; Trueblood and Seibel, 2013). Migratory krill (Euphausia eximia and Nematoscelis gracilis) from this same OMZ region also employ MRD at this PwO2 (Seibel, 2011; Seibel et al., 2016). These are different THRs than those employed by tidepool sculpins and killifish despite similar environmental O2 characteristics, and the reason may involve predation risk. As discussed, predation risk in tidepools and estuaries is high, particularly for small animals in MRD. But predation risk in the OMZ is relatively low owing to low light and activity levels, a diffuse distribution of animals (see Childress, 1995; Drazen and Seibel, 2007; Seibel and Drazen, 2007; Seibel et al., 2000), and low O2 levels that tend to keep top ocean predators such as sharks, tunas and billfishes out (Brill, 1994; Nasby-Lucas et al., 2009; Vetter et al., 2008). Consequently, animals living in the OMZ – particularly those that migrate into oxygenated surface waters to complete necessary behaviours such as feeding and mating – can employ MRD with a relatively low risk of being eaten.

Oceanic OMZs – permanent residents

Permanent OMZ residents experience hypoxia that is moderate (PwO2) and chronic (see Box 2). These animals, which include many fish and invertebrate species, tend to live towards the OMZ's periphery, where PwO2 levels are not as severe as in its centre (Childress and Seibel, 1998). Probably owing to the detrimental effects of chronic reliance on anaerobic glycolysis and/or MRD, these animals rely primarily on aerobic metabolism through a suite of highly effective O2 extraction adaptations.

From a THR perspective, by far the best-studied permanent OMZ resident is the giant red mysid (Gnathophausia ingens). The red mysid has a high ventilatory capacity (Childress, 1971), a large mass-specific gill surface area (Childress, 1971), a small blood–water diffusion distance across the gills (Seibel, 2011), a high circulatory capacity (Belman and Childress, 1976), and a haemocyanin with an extremely high affinity for O2 and a large Bohr effect (see Glossary) to facilitate tissue O2 delivery (Sanders and Childress, 1990a,b). Combined with an extremely low routine metabolic rate (a common trait of permanent OMZ residents; see Childress, 1995), this results in a Pcrit value of 0.8 kPa (Seibel, 2011), coincident with the minimum PwO2 that the mysid typically encounters in the OMZ (Childress and Seibel, 1998). In fact, across a wide range of OMZ residents, Pcrit has been shown to correlate at near-unity (or below) with the minimum PwO2 each of these animals experience in the wild (Childress, 1975; Cowles et al., 1991; Donnelly and Torres, 1988; Torres et al., 1994).

Highly developed mechanisms of O2 extraction and delivery are present in a diverse array of permanent OMZ resident species beyond the giant red mysid (Childress and Seibel, 1998; Lamont and Gage, 2000; Levin, 2003), and these mechanisms may preclude a significant reliance on anaerobic glycolysis. It is believed that anaerobic glycolysis is used by permanent OMZ residents in the same way it is used by species from normoxic habitats, not to support routine metabolic rate, but to supplement supra-routine metabolic rates (Childress and Seibel, 1998; Levin, 2003). Consistent with this, the capacities for anaerobic metabolism (as indicated by maximal rates of anaerobic enzymes) of the permanent OMZ residents that have been investigated are no higher than those of closely related species (or conspecifics) from outside the OMZ (Childress and Somero, 1979; Yang and Somero, 1993; Vetter et al., 2008; Childress and Seibel, 1998; Friedman et al., 2012). Exceptions exist, however. Some permanent OMZ residents such as the copepod Gaussia princeps (Childress, 1976) and the isopod Anuropus bathypelagicus (Childress, 1975) have been caught at PwO2 levels that are lower than their measured Pcrit values (0.8 kPa for G. princeps, 0.6 kPa for A. bathypelagicus) and they may therefore use anaerobic glycolysis to support routine metabolic rates. But such examples are rare (see Childress and Seibel, 1998).

In summary, permanent OMZ residents use a predominantly aerobic THR to survive in their chronically hypoxic habitat. Anaerobic glycolysis is typically reserved for supplementing supra-routine metabolic rates, while MRD has not been measured in these species. This THR enables these species to carry out routine behaviours and life-history events without accumulating a significant O2 debt, despite living permanently in hypoxia. Predominantly aerobic THRs may be selected for in other chronic hypoxic environments of moderate PwO2, such as high-altitude (<3000 m) lakes. The single physiology study on species from these environments that we know of revealed a highly plastic gill surface area in Lake Qinghai (3205 m) naked carp (Matey et al., 2008). However, more work is needed to say anything general about hypoxic adaptations in high-altitude fishes.

Oligotrophic winterfreeze lakes

Oligotrophic (see Glossary) winterfreeze lakes experience hypoxia that is moderately severe (PwO2) and long in duration (see Box 2). A recent study examined the THRs of two threespine stickleback (Gasterosteus aculeatus) populations from two isolated lakes in British Columbia: Alta Lake, which experiences long-term hypoxia owing to winterfreeze, and Trout Lake, which does not (Regan et al., 2017b). The Alta Lake fish were found to be significantly more hypoxia-tolerant than the Trout Lake fish, and although Pcrit and lactate accumulation did not differ, the Alta Lake fish used MRD at sub-PcritPwO2 levels and the Trout Lake fish did not. Interestingly, aspects of the Alta Lake fish's MRD were different than those of another MRD-inducing species, the goldfish (discussed in detail below). Alta Lake sticklebacks depress metabolic rate by 33% and do so at 2.8 kPa PwO2, whereas goldfish depress metabolic rate by 80% and wait until near-anoxia to initiate it. These MRD differences may relate to variation in each species' hypoxic environment. Although the native lakes of goldfish likely become anoxic during wintertime (like the native lakes of crucian carp; Vornanen, 2004), apart from at the water–sediment interface (Dunnington et al., 2016), Alta Lake waters do not reach anoxia (Jacques Whitford AXYS Ltd, 2007). Selection may therefore be acting on hypoxic survival strategies at higher PwO2 values in the Alta Lake environment than in the more severe goldfish environment.

Eutrophic winterfreeze lakes

Eutrophic (see Glossary) winterfreeze lakes are typified by hypoxia that is severe (PwO2) and long in duration (see Box 2). Consequently, they tend to be colonized by highly tolerant fishes, such as the well-studied crucian carp (Carassius carassius; Vornanen et al., 2009) and its congener, the goldfish (Carassius auratus; Ultsch, 1989). These fishes employ a complex THR that involves highly effective mechanisms of aerobic metabolism, anaerobic metabolism and MRD. Aspects of this THR are altered not only in response to PwO2, rate of hypoxia induction and duration of hypoxia exposure, but also in anticipation of the naturally occurring hypoxic season. Despite this THR's complexity, the evidence suggests it is altered so as to maintain routine metabolic rate in as many hypoxic environments as possible. This is accomplished in different ways depending on time: for rapid induction rates and acute exposure durations (−21 kPa PwO2 h−1, 1 h exposure), goldfish upregulate glycolysis to buffer ATP supply so as to maintain routine metabolic rate (Regan et al., 2017a); whereas for gradual induction rates and long exposure durations (−2.6 kPa PwO2 h−1, 8 h exposure), goldfish and carp increase environmental O2 extraction by increasing gill surface area and Hb–O2 affinity so as to support routine metabolic rate aerobically (Sollid et al., 2003; Tzaneva et al., 2011; Dhillon et al., 2013; Regan and Richards, 2017). The combination of a highly plastic gill surface area, the highest Hb–O2 binding affinities reported for vertebrates (Burggren, 1982; Sollid et al., 2003; Regan et al., 2017a), and a generally low routine metabolic demand allows goldfish to maintain routine metabolic rates down to ∼1 kPa PwO2.

At PwO2 below ∼1 kPa, goldfish rapidly depress metabolic rate to ∼20% of routine values in less than 20 min (van Ginneken and van den Thillart, 2009; Regan et al., 2013). O2 content in the water is negligible or altogether absent at these PwO2 levels, and the goldfish and carp become solely reliant on glycolysis to supply the ATP required to fuel their reduced metabolic demands. The use of MRD and glycolysis initially leads to an accumulation of lactate and protons (Regan et al., 2017a), but as the hypoxic/anoxic exposure duration lengthens, the fish begin to convert these end-products into ethanol, which they excrete across their gills and thus mitigate a metabolic acidosis (Shoubridge and Hochachka, 1980; Holopainen et al., 1986; Regan et al., 2017a). The glycogen stores of goldfish and crucian carp are larger than those of any other fish species (Richards, 2009), and, similar to fat stores in hibernating mammals, these stores significantly increase in size over the late summer and early autumn to levels that are sufficient to fuel the fish's depressed metabolic rates in anoxia during the winter months (Vornanen et al., 2009).

Another species system native to eutrophic winterfreeze lakes is the centrarchid sunfishes. The ranges of two closely related sunfish, the bluegill (Lepomis macrochirus) and pumpkinseed (Lepomis gibbosus), overlap, but the most northern lakes – the ones that experience the most severe winterfreeze hypoxia – contain only pumpkinseed (Mittelbach, 1984; Farwell et al., 2007). Unsurprisingly, pumpkinseed have repeatedly been shown to be more hypoxia-tolerant than bluegill (Farwell et al., 2007; Mathers et al., 2014; Borowiec et al., 2016). The pumpkinseed's greater tolerance does not appear to result from greater aerobic abilities, as the two species do not differ in Pcrit (Mathers et al., 2014; Borowiec et al., 2016) or a wide array of underlying mechanisms (Crans et al., 2015; Borowiec et al., 2016). Where the species do differ is in their capacities for anaerobic metabolism, with the more tolerant pumpkinseed displaying higher lactate dehydrogenase activities in axial muscle (Farwell et al., 2007; Borowiec et al., 2016) and heart (Borowiec et al., 2016) than the less tolerant bluegill. The pumpkinseed's greater anaerobic capacity may contribute to its greater hypoxia tolerance, but because anaerobic glycolysis on its own is a limited long-term strategy owing to glycogen depletion and end-product accumulation, it is unlikely to solely explain how pumpkinseed can tolerate the more hypoxic northern lakes that bluegill cannot. It may be that pumpkinseed, like goldfish, rely on MRD and/or plastic mechanisms that enhance O2 extraction, and that their capacities for these traits are greater than the bluegill's. This warrants further investigation.

Swamps

The swamps surrounding Lake Victoria experience hypoxia that is chronic and severe in PwO2 (see Box 2). The haplochromine cichlids that live here rely on a primarily aerobic THR. Compared with closely related lake-dwelling (i.e. normoxic) species or conspecifics, the swamp-dwelling fish have larger gill surface areas, higher haematocrits and Hb concentrations, and lower routine O2 demands. These modifications result in swamp-dwelling fishes exhibiting extremely low Pcrit values that are approximately half those of lake-dwelling fishes and just slightly higher than the lowest PwO2 measured in these habitats (Chapman et al., 2002). Their hypoxia tolerance is further enhanced by effective ASR abilities, which are a central component of their overall THR (Chapman et al., 1995). Moreover, swamp dwellers engage ASR at significantly lower PwO2 than do lake dwellers (which are also capable of ASR; Chapman et al., 2002), a possible advantage given the high daytime predation pressures exerted by pied kingfishers (Ceryle rudis) (Randle and Chapman, 2004). The swamp dwellers' low Pcrit values enable them to remain in deeper, safer waters throughout the day when the kingfisher is active and when photosynthetic activity elevates PwO2 above Pcrit.

It is not known whether these fishes use MRD, but it is known that the swamp dwellers' capacities for anaerobic metabolism are no higher than those of closely related lake-dwelling fishes (as indicated by anaerobic enzyme activities in various tissues of Pseudocrenilabrus multicolor populations from swamp and lake habitats; Crocker et al., 2013). It therefore appears that swamp dwellers employ a predominantly aerobic THR, though additional MRD-focused work is needed to provide a complete picture of their THR.

Amazonian fishes can experience hypoxia that is mild or severe in PwO2, and short or long in duration (see Box 2). Perhaps because of the diversity and widespread persistence of Amazonian hypoxic habitats, many Amazonian fish species have independently evolved high tolerances to hypoxia (Almeida-Val and Val, 1993). Most of these species achieve this through behavioural and/or morphological features that maximize their abilities to acquire O2 in their O2-depleted habitats, the prime examples being ASR and air breathing. Many Amazonian fish families have independently evolved morphological features to optimize ASR and air breathing, including extensible lower lips to syphon O2-rich water directly across the gills, and a wide variety of air-breathing organs ranging from modified buccal cavities to lungs (Val et al., 1998). These morphological features, and the behaviours they optimize, are believed to have evolved in response to aquatic hypoxia (Graham, 1997; Kramer and McClure, 1982). It is no surprise then that they are widely used among the Amazon's hypoxia-dwelling fish species.

One study collected the resident species from an isolated Amazonian lake (Camaleao Lake) after it had become severely hypoxic, and then determined the primary hypoxia tolerance strategy used by each species. Of the 11 families caught (numerous species for most), seven used air-breathing as their primary means of tolerating hypoxia, two used ASR, one used Hb–O2 affinity modulation and one used MRD (Junk et al., 1983). In a similar study in which 20 species were caught in a hypoxic Amazonian lake, 10 species used ASR as a primary means of tolerating hypoxia, four used air-breathing, four positioned themselves directly adjacent to O2-secreting plant roots, one combined a large gill surface area with a high Hb–O2 affinity and one used MRD (the Amazonian oscar, Astronotus ocellatus, the most tolerant of the group; Soares et al., 2006). Furthermore, although air breathing and/or ASR behaviours increase the susceptibility to aerial predation (Kramer and Mehegan, 1981), some of these fishes have evolved complex group behaviours to mitigate this risk (Sloman et al., 2009).

In addition to ASR and air breathing, hypoxia-adapted Amazonian fishes tend to possess characteristics across the multiple steps of the O2 cascade that enhance O2 extraction and delivery. These include high ventilation rates, large gill surface areas (Saint-Paul, 1984), and high blood–O2 carrying capacities through increased red blood cell count, [Hb] and Hb–O2 binding affinities (Saint-Paul, 1984; Val and Almeida-Val, 1995; Muusze et al., 1998; Val et al., 1998; Affonso et al., 2002). Beyond sustaining aerobic metabolism, hypoxia-exposed Amazonian fishes also strongly activate anaerobic metabolism (e.g. glycolysis, CrP hydrolysis) (Chippari-Gomes et al., 2005; MacCormack et al., 2006; Richards et al., 2007; Scott et al., 2008). And at least one species – the Amazonian oscar – uses MRD, evidenced by suppressed O2 and ATP-consuming processes and enzymes, such as protein synthesis and Na+/K+-ATPase, respectively (Muusze et al., 1998; Lewis et al., 2007; Richards et al., 2007; Scott et al., 2008). Given the Amazon's species diversity and hypoxic heterogeneity, it is perhaps unsurprising that a wide range of THRs would be seen among Amazonian fishes. However, as not all aspects of the THR are available for all species discussed (e.g. scarcity of MRD information), it is also possible that THRs of Amazonian fishes are more similar than the data currently suggest.

We can combine the 10 case studies summarized above with the environmental hypoxia matrix in Fig. 2 to draw some general conclusions on how the THR may relate to the hypoxic environment (Fig. 3). Importantly, these are generalizations based on the limited number of available studies, and more specifically, the relatively small number explored here. As additional studies are completed on species native to environments within each hypoxia matrix quadrant, these generalizations will become more refined and likely modified in various ways. For now, however, we will use currently available information from across studies to build an initial environment-focused THR framework. The individual studies we used to do so (which are highlighted in Table S1) were chosen based on two main criteria: (i) they used sound methods to measure mechanisms that are directly related to one or more THR metabolic modes; and (ii) they were carried out on species native to hypoxic environments that are well characterized.

Fig. 3.

The total hypoxic responses (THRs) of species inhabiting the different hypoxic quadrants. The circular pie charts represent THR. Each pie chart's total area qualitatively represents its total metabolic rate relative to the size of the pie chart in Q1, which represents normoxic routine metabolic rate. A smaller-diameter pie chart therefore represents a depressed metabolic rate. The black-filled portion of a pie chart qualitatively represents the aerobic contribution to that THR's metabolic rate, and the white-filled portion represents the anaerobic contribution (ratios are approximate). All-black and all-white pie charts therefore represent fully aerobic and anaerobic THRs, respectively. ‘Access to higher O2’ labels indicate environments in which the inhabitants may access higher PwO2 regions within their hypoxic environment (e.g. via aerial emergence).

Fig. 3.

The total hypoxic responses (THRs) of species inhabiting the different hypoxic quadrants. The circular pie charts represent THR. Each pie chart's total area qualitatively represents its total metabolic rate relative to the size of the pie chart in Q1, which represents normoxic routine metabolic rate. A smaller-diameter pie chart therefore represents a depressed metabolic rate. The black-filled portion of a pie chart qualitatively represents the aerobic contribution to that THR's metabolic rate, and the white-filled portion represents the anaerobic contribution (ratios are approximate). All-black and all-white pie charts therefore represent fully aerobic and anaerobic THRs, respectively. ‘Access to higher O2’ labels indicate environments in which the inhabitants may access higher PwO2 regions within their hypoxic environment (e.g. via aerial emergence).

For hypoxia exposures that are short (time) and moderate (PwO2) (Q1), we believe the THR may vary with the ability to escape to more oxygenated waters. If escape to more oxygenated waters is not possible, the fishes tend to employ a primarily aerobic THR with some contribution of anaerobic metabolism. Generally, Q1 fishes are hypoxia intolerant relative to Q2, Q3 and Q4 species.

For hypoxia exposures that are short and severe (Q2), we believe the THR may be heavily influenced by aerial/surface access and predators (Fig. 3); if the air–water interface is accessible (e.g. tidepools), fishes living in these environments tend to prioritize aerobic metabolism by using ASR and/or air breathing. If an aerial predator presents itself, fishes tend to dive into the hypoxic water and buffer routine metabolic rate using anaerobic glycolysis, which in this environment may be more practical than inducing MRD because it allows the fishes to maintain cellular energy balance (though not indefinitely) without impairing predator avoidance behaviour as a result of a metabolic shut-down. The costs of anaerobic metabolism (fuel depletion, end-product accumulation) are positively correlated with time spent in the hypoxic environment, and, for these fishes, this would typically be short: either the predation threat will subside and allow the fish to re-emerge and/or perform ASR, or the O2 will be replenished by photosynthesis and/or the rising tide. The costs accrued with anaerobic reliance would therefore be low. For Q2 environments that lack an air–water interface, the response tends to depend on PwO2. If PwO2 is above residents' Pcrit (e.g. less hypoxic OMZ regions), then aerobic metabolism will be prioritized. However, if PwO2 is below residents' Pcrit (e.g. more hypoxic OMZ regions), then aerobic metabolism cannot be sustained. Animals living in these environments tend to use MRD over anaerobic metabolism, perhaps enabled by relatively low predation pressures. Deep MRD is used if predation risk is low to absent (e.g. jumbo squid), and moderate MRD is used if predation risk is moderate to low (e.g. krill).

For exposures that are long and moderate (Q3), we believe the THR may vary as a function of exposure duration (Fig. 3). Species that live in chronically moderate hypoxic environments (e.g. OMZ periphery, high-altitude lakes) tend to rely on enhanced O2 extraction abilities to support metabolic rate aerobically. Because these species infrequently enter fully oxygenated waters, they need to support routine activities such as feeding and mating in chronic hypoxia, and this makes an aerobically based THR ideal. Species that live in seasonally moderate hypoxic environments (e.g. oligotrophic winterfreeze lakes) tend to use modest MRD, perhaps because energetically costly routine activities such as mating and reproduction are typically accomplished during the oxygenated months of the year. More work is needed on species from seasonally fluctuating moderately hypoxic environments.

Finally, for exposures that are long and severe (Q4), we believe the THR may vary as a function of aerial/surface access (Fig. 3). If it is accessible (e.g. Amazon basin), then the fishes living in these environments tend to exploit its high O2 content using ASR and/or air breathing, effectively uncoupling themselves from their severely hypoxic aquatic environment. If the air–water interface is not available (e.g. eutrophic winterfreeze lakes), then aerobic metabolism is not an option and the fishes living here tend to employ deep MRD so as to conserve limited anaerobic fuel reserves. A general lack of predators in these environments allows these fishes to surrender locomotor performance with minimal predation-related consequences, while the return of O2 with spring thaw allows them to complete routine activities such as feeding and mating in oxygenated waters.

Looking across these studies as a function of hypoxic environmental variation, it is apparent that the hypoxic environment is not the sole influence on THR. It is possible that other factors such as genetic constraint, rate of environmental change (e.g. climate change-related) and complex interactions with other abiotic factors (e.g. temperature, PwCO2, pH) may also contribute to a species' THR, and divorcing these effects from those of the hypoxic environment is difficult. However, there appear to be common patterns in the THRs across species that depend on their native hypoxic environments. Using the extensively studied Q2, we illustrate that distantly related species share similar THRs, suggesting that convergent evolution has played a role in shaping the hypoxic metabolic phenotype. Spanning different phylogenetic orders, sculpins (Scorpaeniformes), triplefins (Perciformes), killifish (Cyprinodontiformes) and the plainfin midshipman (Betrachoidiformes) are commonly found to have an enhancement of aerobic metabolism along with contributions from anaerobic metabolism in defense of hypoxic stress (although the scarcity of direct MRD measurements in ecologically relevant time frames should be noted; Fig. 4). The THR is more similar in the distantly related species than with the species' respective taxonomic counterparts. For example, hypoxia-tolerant sculpin and triplefin species inhabiting the intertidal zone share more similar traits with each other, e.g. low Pcrit and high Hb–O2 affinity, than with the less tolerant subtidal sculpin and triplefin species, respectively, e.g. high Pcrit and low Hb–O2 affinity (Mandic et al., 2009a; Brix et al., 1999). This would suggest a convergence of similar metabolic strategies among distantly related species. However, disentangling the roles of phylogenetic history and the hypoxic environment per se on shaping species' THRs will require a more widespread use of rigorous, phylogenetically appropriate comparisons.

Fig. 4.

Phylogeny and the traits underlying the THR of representative fishes from Q2. A truncated phylogeny of well-studied examples of Q2 fishes – sculpins (order Scorpaeniformes), triplefins (order Perciformes), killifish (Cyprinodontiformes) and plainfin midshipman (order Batrachoidiformes) – is represented (modified from Nelson, 2006). Listed are the known traits underlying the THR of each group of species.

Fig. 4.

Phylogeny and the traits underlying the THR of representative fishes from Q2. A truncated phylogeny of well-studied examples of Q2 fishes – sculpins (order Scorpaeniformes), triplefins (order Perciformes), killifish (Cyprinodontiformes) and plainfin midshipman (order Batrachoidiformes) – is represented (modified from Nelson, 2006). Listed are the known traits underlying the THR of each group of species.

As current human practices increase the prevalence and severity of hypoxia among the world's aquatic environments (IPCC, 2014; Schmidtko et al., 2017; Smith et al., 2006), understanding how metabolic strategies associate with different hypoxic environments is probably more important than ever. Knowing which THRs are most conducive to survival in which types of hypoxic environments may help us to better identify potentially vulnerable species, and better predict their redistribution patterns as their environments become increasingly hypoxic.

We thank Drs Rush Dhillon and Alex Zimmer, as well as David Hatton and the two anonymous reviewers for their insightful comments.

Author contributions

M.M. and M.D.R. co-developed the ideas and co-wrote the manuscript.

Funding

M.M. was supported by a Natural Sciences and Engineering Research Council of Canada postdoctoral fellowship.

Abbaraju
,
N. V.
and
Rees
,
B. B.
(
2011
).
Effects of dissolved oxygen on glycolytic enzyme specific activities in liver and skeletal muscle of Fundulus heteroclitus
.
Fish Physiol. Biochem.
38
,
615
-
624
.
Affonso
,
E. G.
,
Polez
,
V. P.
,
Correa
,
C. F.
,
Mazon
,
A. F.
,
Araujo
,
M.
,
Moraes
,
G.
and
Rantin
,
F. T.
(
2002
).
Blood parameters and metabolites in the teleost fish Colossoma macropomum exposed to sulfide or hypoxia
.
Comp. Biochem. Phys. C
133
,
375
-
382
.
Albert
,
J. S.
and
Reis
,
R.
(
2011
).
Historical Biogeography of Neotropical Freshwater Fishes
.
Berkeley, CA
:
University of California Press
.
Almeida-Val
,
V.
and
Val
,
A. L.
(
1993
).
Hypoxia tolerance in Amazon fishes: status of an unde-explored biological ‘goldmine
’. In
Surviving Hypoxia: Mechanisms of Control and Adaptation
(ed.
P.
Hochachka
,
P.
Lutz
,
T. J.
Sick
and
M.
Rosenthal
), pp.
435
-
445
.
Boca Raton, FL
:
CRC Press
.
Barica
,
J.
and
Mathias
,
J. A.
(
1979
).
Oxygen depletion and winterkill risk in small prairie lakes under extended ice cover
.
J. Fish. Res. Board Can.
36
,
980
-
986
.
Belman
,
B. W.
and
Childress
,
J. J.
(
1976
).
Circulatory adaptations to the oxygen minimum layer in the bathypelagic mysid Gnathophausia ingens
.
Biol. Bull.
150
,
15
-
37
.
Borowiec
,
B. G.
,
Darcy
,
K. L.
,
Gillette
,
D. M.
and
Scott
,
G. R.
(
2015
).
Distinct physiological strategies are used to cope with constant hypoxia and intermittent hypoxia in killifish (Fundulus heteroclitus)
.
J. Exp. Biol.
218
,
1198
-
1211
.
Borowiec
,
B. G.
,
Crans
,
K. D.
,
Khajali
,
F.
,
Pranckevicius
,
N. A.
,
Young
,
A.
and
Scott
,
G. R.
(
2016
).
Interspecific and environment-induced variation in hypoxia tolerance in sunfish
.
Comp. Biochem. Physiol. A Mol. Integr. Physiol.
198
,
59
-
71
.
Breitburg
,
D. L.
(
1990
).
Near-shore hypoxia in the Chesapeake Bay: patterns and relationships among physical factors
.
Estuarine Coastal Shelf Sci.
30
,
593
-
609
.
Brett
,
J. R.
and
Groves
,
T. D. D.
(
1979
).
Physiological energetics
. In
Fish Physiology
, Vol.
8
(ed.
W. S.
Hoar
,
D. J.
Randall
and
J. R.
Brett
), pp.
280
-
352
.
London
:
Elsevier
.
Bridges
,
C. R.
(
1988
).
Respiratory adaptations in intertidal fish
.
Am. Zool.
28
,
79
-
96
.
Brill
,
R. W.
(
1994
).
A review of temperature and oxygen tolerance studies of tunas pertinent to fisheries oceanography, movement models and stock assessments
.
Fish. Oceanogr.
3
,
204
-
216
.
Brix
,
O.
,
Clements
,
K. D.
and
Wells
,
R. M. G.
(
1999
).
Haemoglobin components and oxygen transport in relation to habitat distribution in triplefin fishes (Tripterygiidae)
.
J. Comp. Physiol. B.
169
,
329
-
334
.
Burggren
,
W. W.
(
1982
).
“Air gulping” improves blood oxygen transport during aquatic hypoxia in the goldfish, Carassius auratus
.
Physiol. Zool.
55
,
327
-
334
.
Burggren
,
W.
and
Roberts
,
J.
(
1991
).
Respiration and metabolism
. In
Environmental and Metabolic Animal Physiology: Comparative Animal Physiology
(ed.
C. L.
Prosser
), pp.
353
-
435
.
New York
:
Wiley-Liss
.
Burnett
,
K. G.
,
Bain
,
L. J.
,
Baldwin
,
W. S.
and
Callard
,
G. V.
(
2007
).
Fundulus as the premier teleost model in environmental biology: opportunities for new insights using genomics
.
Comp. Biochem. Physiol. D
2
,
257
-
286
.
Burton
,
R. S.
and
Reichman
,
O. J.
(
1999
).
Does immune challenge affect torpor duration?
Funct. Ecol.
13
,
232
-
237
.
Carey
,
H. V.
,
Frank
,
C. L.
and
Seifert
,
J. P.
(
2000
).
Hibernation induces oxidative stress and activation of NF-κB in ground squirrel intestine
.
J. Comp. Physiol. B
170
,
551
-
559
.
Chapman
,
L. J.
,
Kaufman
,
L. S.
and
Chapman
,
C. A.
(
1995
).
Hypoxia tolerance in twelve species of East African cichlids: potential for low oxygen refugia in Lake Victoria
.
Conserv. Biol.
9
,
1274
-
1288
.
Chapman
,
L. J.
,
Chapman
,
C. A.
,
Nordlie
,
F. G.
and
Rosenberger
,
A. E.
(
2002
).
Physiological refugia: swamps, hypoxia tolerance and maintenance of fish diversity in the Lake Victoria region
.
Comp. Biochem. Physiol. A
133
,
421
-
437
.
Chevin
,
L.-M.
,
Martin
,
G.
and
Lenormand
,
T.
(
2010
).
Fisher's model and the genomics of adaptation: restricted pleiotropy, heterogenous mutation, and parallel evolution
.
Evolution
64
,
3213
-
3231
.
Chiba
,
K.
(
1983
).
The effect of dissolved oxygen on the growth of young silver bream (Sparus sarba)
.
Bull. Jap. Soc. Sci. Fish.
49
,
601
-
610
.
Childress
,
J. J.
(
1971
).
Respiratory adaptations to the oxygen minimum layer in the bathypelagic mysid Gnathophausia ingens
.
Biol. Bull.
141
,
109
-
121
.
Childress
,
J. J.
(
1975
).
The respiratory rates of midwater crustaceans as a function of depth of occurrence and relation to the oxygen minimum layer off Southern California
.
Comp. Biochem. Physiol. A
50
,
787
-
799
.
Childress
,
J. J.
(
1976
).
Effects of pressure, temperature and oxygen on the oxygen consumption rate of the midwater copepod Gaussia princeps
.
Mar. Biol.
39
,
19
-
24
.
Childress
,
J. J.
(
1995
).
Are there physiological and biochemical adaptations of metabolism in deep-sea animals?
Trends Ecol. Evol.
10
,
30
-
36
.
Childress
,
J. J.
and
Seibel
,
B. A.
(
1998
).
Life at stable low oxygen levels: adaptations of animals to oceanic oxygen minimum layers
.
J. Exp. Biol.
201
,
1223
-
1232
.
Childress
,
J. J.
and
Somero
,
G. N.
(
1979
).
Depth-related enzymic activities in muscle, brain and heart of deep-living pelagic marine teleosts
.
Mar. Biol.
52
,
273
-
283
.
Chippari-Gomes
,
A. R.
,
Gomes
,
L. C.
,
Lopes
,
N. P.
,
Val
,
A. L.
and
Almeida-Val
,
V. M. F.
(
2005
).
Metabolic adjustments in two Amazonian cichlids exposed to hypoxia and anoxia
.
Comp. Biochem. Physiol. B
141
,
347
-
355
.
Christin
,
P.-A.
,
Weinreich
,
D. M.
and
Besnard
,
G.
(
2010
).
Causes and evolutionary significance of genetic convergence
.
Trends Genet.
26
,
400
-
405
.
Cochran
,
R. E.
and
Burnett
,
L. E.
(
1996
).
Respiratory responses of the salt marsh animals, Fundulus heteroclitus, Leiostomus xanthurus, and Palaemonetes pugio to environmental hypoxia and hypercapnia and to the organophosphate pesticide, azinphosmethyl
.
J. Exp. Mar. Biol. Ecol.
195
,
125
-
144
.
Congleton
,
J. L.
(
1980
).
Observations on the responses of some southern California tidepool fishes to nocturnal hypoxic stress
.
Comp. Biochem. Physiol. A
66
,
719
-
722
.
Conte
,
G. L.
,
Arnegard
,
M. E.
,
Peichel
,
C. L.
and
Schluter
,
D.
(
2012
).
The probability of genetic parallelism and convergence in natural populations
.
Proc. R. Soc. B
279
,
5039
-
5047
.
Cowles
,
D. L.
,
Childress
,
J. J.
and
Wells
,
M. E.
(
1991
).
Metabolic rates of midwater crustaceans as a function of depth of occurrence off the Hawaiian Islands: food availability as a selective factor?
Mar. Biol.
110
,
75
-
83
.
Craig
,
P. M.
,
Fitzpatrick
,
J. L.
,
Walsh
,
P. J.
,
Wood
,
C. M.
and
McClelland
,
G. B.
(
2014
).
Coping with aquatic hypoxia: how the plainfin midshipman (Porichthys notatus) tolerates the intertidal zone
.
Environ. Biol. Fish.
97
,
163
-
172
.
Crans
,
K. D.
,
Pranckevicius
,
N. A.
and
Scott
,
G. R.
(
2015
).
Physiological tradeoffs may underlie the evolution of hypoxia tolerance and exercise performance in sunfish (Centrarchidae)
.
J. Exp. Biol.
218
,
3264
-
3275
.
Crocker
,
C. D.
,
Chapman
,
L. J.
and
Martínez
,
M. L.
(
2013
).
Hypoxia-induced plasticity in the metabolic response of a widespread cichlid
.
Comp. Biochem. Physiol. B. Biochem. Mol. Biol.
166
,
141
-
147
.
D'Avanzo
,
C.
and
Kremer
,
J. N.
(
1994
).
Diel oxygen dynamics and anoxic events in an eutrophic estuary of Waquoit Bay, Massachusetts
.
Estuaries
17
,
131
-
139
.
Daxboeck
,
C.
and
Heming
,
T. A.
(
1982
).
Bimodal respiration in the intertidal fish, Xiphister atropurpureus (Kittlitz)
.
Mar. Freshw. Behav. Phys.
9
,
23
-
33
.
Dhillon
,
R. S.
,
Yao
,
L.
,
Matey
,
V.
,
Chen
,
B.-J.
,
Zhang
,
A.-J.
,
Cao
,
Z.-D.
,
Fu
,
S.-J.
,
Brauner
,
C. J.
,
Wang
,
Y. S.
and
Richards
,
J. G.
(
2013
).
Interspecific differences in hypoxia-induced gill remodeling in carp
.
Physiol. Biochem. Zool.
86
,
727
-
739
.
Donnelly
,
J.
and
Torres
,
J. J.
(
1988
).
Oxygen consumption of midwater fishes and crustaceans from the eastern Gulf of Mexico
.
Mar. Biol.
97
,
483
-
494
.
Doudoroff
,
P.
and
Shumway
,
D. L.
(
1970
).
Dissolved oxygen requirements of freshwater fishes
.
FAO Fisheries Technical Paper No. 86
.
Rome: FAO
.
Dowd
,
W. W.
,
Renshaw
,
G. M. C.
,
Cech
,
J. J.
Jr
, and
Kültz
,
D.
(
2010
).
Compensatory proteome adjustments imply tissue-specific structural and metabolic reorganization following episodic hypoxia or anoxia in the epaulette shark (Hemiscyllium ocellatum)
.
Physiol. Genomics
42
,
93
-
114
.
Draud
,
M.
,
Bossert
,
M.
and
Zimnavoda
,
S.
(
2004
).
Predation on hatchling and juvenile diamondback terrapins (Malaclemys terrapin) by the Norway rat (Rattus norvegicus)
.
J. Herpetol.
38
,
467
-
470
.
Drazen
,
J. C.
and
Seibel
,
B. A.
(
2007
).
Depth-related trends in metabolism of benthic and benthopelagic deep-sea fishes
.
Limnol. Oceanogr.
52
,
2306
-
2316
.
Driedzic
,
W. R.
and
Gesser
,
H.
(
1994
).
Energy metabolism and contractility in ectothermic vertebrate hearts: hypoxia, acidosis and low temperature
.
Physiol. Rev.
74
,
221
-
259
.
Dunn
,
J. F.
and
Hochachka
,
P. W.
(
1986
).
Metabolic responses of trout (Salmo gairdneri) to acute environmental hypoxia
.
J. Exp. Biol.
123
,
229
-
242
.
Dunn
,
J. F.
and
Hochachka
,
P. W.
(
1987
).
Turnover rates of glucose and lactate in rainbow trout during acute hypoxia
.
Can. J. Zool.
65
,
1144
-
1148
.
Dunnington
,
D. W.
,
Spooner
,
I. S.
,
White
,
C. E.
,
Cornett
,
R. J.
,
Williamson
,
D.
and
Nelson
,
M.
(
2016
).
A geochemical perspective on the impact of development at Alta Lake, British Columbia, Canada
.
J. Paleolimnol.
56
,
315
-
330
.
Estok
,
P.
,
Zsebok
,
S.
and
Siemers
,
B.
(
2009
).
Great tits search for, capture, kill and eat hibernating bats
.
Biol. Lett
.
6
,
59
-
62
.
Everett
,
M. V.
,
Antal
,
C. E.
and
Crawford
,
D. L.
(
2012
).
The effect of short-term hypoxic exposure on metabolic gene expression
.
J. Exp. Zool.
317A
,
9
-
23
.
Farwell
,
M.
,
Fox
,
M. G.
,
Moyes
,
C. D.
and
Burness
,
G.
(
2007
).
Can hypoxia tolerance explain differences in distribution of two co-occurring north temperate sunfishes?
Environ. Biol. Fish.
78
,
83
-
90
.
Fitzgibbon
,
Q. P.
,
Seymour
,
R. S.
,
Ellis
,
D.
and
Buchanan
,
J.
(
2007
).
The energetic consequence of specific dynamic action in southern bluefin tuna Thunnus maccoyii
.
J. Exp. Biol.
210
,
290
-
298
.
Friedman
,
J. R.
,
Condon
,
N. E.
and
Drazen
,
J. C.
(
2012
).
Gill surface area and metabolic enzyme activities of demersal fishes associated with the oxygen minimum zone off California
.
Limnol. Oceanogr.
57
,
1701
-
1710
.
Gallaugher
,
P.
and
Farrell
,
A. P.
(
1998
).
Hematocrit and blood oxygen-carry capacity
. In
Fish Physiology
, Vol.
17
(ed.
S. F.
Perry
and
B. L.
Tufts
), pp.
185
-
227
.
San Diego, CA
:
Academic Press
.
Goldshmid
,
R.
,
Holzman
,
R.
,
Weihs
,
D.
and
Genin
,
A.
(
2004
).
Aeration of corals by sleep-swimming fish
.
Limnol. Oceanogr.
49
,
1832
-
1839
.
Gonzalez
,
R. R.
and
Quiñones
,
R. A.
(
2002
).
LDH activity in Euphausia mucronata and Calanus chilensis: implications for vertical migration behaviour
.
J. Plankton Res.
24
,
1349
-
1356
.
Goto
,
T.
,
Nishida
,
K.
and
Nakaya
,
K.
(
1999
).
Internal morphology and function of paired fins in the epaulette shark, Hemiscyllium ocellatum
.
Ichthyol. Res.
46
,
281
-
289
.
Gracey
,
A. Y.
,
Troll
,
J. V.
and
Somero
,
G. N.
(
2001
).
Hypoxia-induced gene expression profiling in the euryoxic fish Gillichthys mirabilis
.
Proc. Natl. Acad. Sci. USA
98
,
1993
-
1998
.
Graham
,
J. B.
(
1997
).
Air-Breathing Fishes: Evolution, Diversity, and Adaptation
.
San Diego, CA
:
Academic Press
.
Hickey
,
A. J. R.
,
Renshaw
,
G. M. C.
,
Speers-Roesch
,
B.
,
Richards
,
J. G.
,
Wang
,
Y.
,
Farrell
,
A. P.
and
Brauner
,
C. J.
(
2012
).
A radical approach to beating hypoxia: depressed free radical release from heart fibres of the hypoxia-tolerant epaulette shark (Hemiscyllum ocellatum)
.
J. Comp. Physiol. B
182
,
91
-
100
.
Hill
,
J. V.
,
Davison
,
W.
and
Marsden
,
I. D.
(
1996
).
Aspects of the respiratory biology of two New Zealand intertidal fishes, Acanthoclinus fuscus and Forsterygion sp
.
Environ. Biol. Fish.
45
,
85
-
93
.
Hilton
,
Z.
,
Wellenreuther
,
M.
and
Clements
,
K. D.
(
2008
).
Physiology underpins habitat partitioning in a sympatric sister-species pair of intertidal fishes
.
Funct. Ecol.
22
,
1108
-
1117
.
Hochachka
,
P. W.
and
Somero
,
G. N.
(
2002
).
Biochemical Adaptation: Mechanism and Process in Physiological Evolution
.
New York
:
Oxford University Press
.
Hochachka
,
P. W.
,
Buck
,
L. T.
,
Doll
,
C. J.
and
Land
,
S. C.
(
1996
).
Unifying theory of hypoxia tolerance: molecular/metabolic defense and rescue mechanisms for surviving oxygen lack
.
Proc. Natl. Acad. Sci. USA
93
,
9493
-
9498
.
Holeton
,
G. F.
and
Randall
,
D. J.
(
1967a
).
Changes in blood pressure in the rainbow trout during hypoxia
.
J. Exp. Biol.
46
,
297
-
305
.
Holeton
,
G. F.
and
Randall
,
D. J.
(
1967b
).
The effect of hypoxia upon the partial pressure of gases in the blood and water afferent and efferent to the gills of rainbow trout
.
J. Exp. Biol.
46
,
317
-
327
.
Holopainen
,
I. J.
,
Hyvärinen
,
H.
and
Piironen
,
J.
(
1986
).
Anaerobic wintering of crucian carp (Carassius carassius L.) II. Metabolic products
.
Comp. Biochem. Physiol.
83
,
239
-
242
.
Hughes
,
G. M.
and
Saunders
,
R. L.
(
1970
).
Responses of the respiratory pumps to hypoxia in the rainbow trout (Salmo gairdneri)
.
J. Exp. Biol.
53
,
529
-
545
.
Hugie
,
D. M.
,
Thuringer
,
P. L.
and
Smith
,
R. J. F.
(
1991
).
The response of the tidepool sculpin, Oligocottus maculosus, to chemical stimuli from injured conspecifics, alarm signalling in the Cottidae (Pisces)
.
Ethology
89
,
322
-
334
.
Humphries
,
M. M.
,
Thomas
,
D. W.
and
Kramer
,
D. L.
(
2003
).
The role of energy availability in Mammalian hibernation: a cost-benefit approach
.
Physiol. Biochem. Zool.
76
,
165
-
179
.
Innes
,
A. J.
and
Wells
,
R. M. G.
(
1985
).
Respiration and oxygen transport functions of the blood from an intertidal fish, Helcogramma medium (Tripterygiidae)
.
Environ. Biol. Fish.
14
,
213
-
226
.
IPCC
(
2014
).
Climate Change 2014. Impacts, Adaptation, and Vulnerability. Part B: Regional Aspects. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change (ed. V. R. Barros, C. B. Field, D. J. Dokken, M. D. Mastrandrea, K. J. Mach, T. E. Bilir, M. Chatterjee, K. L. Ebi, Y. O. Estrada, R. C. Genova, et al.). Cambridge University Press.
Itazawa
,
Y.
and
Takeda
,
T.
(
1978
).
Gas exchange in the carp gills in normoxic and hypoxic conditions
.
Resp. Physiol.
35
,
263
-
269
.
Jacques Whitford AXYS Ltd (2007). Alta Lake Limnology Study, Resort Municipality of Whistler. Project No. 1024544.
Johnston
,
I. A.
(
1977
).
A comparative study of glycolysis in red and white muscles of the trout (Salmo gairdneri) and mirror carp (Cyprinus carpio)
.
J. Fish Biol.
11
,
575
-
588
.
Junk
,
W. J.
,
Soares
,
G. M.
and
Carvalho
,
F. M.
(
1983
).
Distribution of fish species in a lake of the Amazon River floodplain near Manaus (lago Camaleao), with special reference to extreme oxygen conditions
.
Amazoniana
7
,
397
-
431
.
Kneib
,
R. T.
(
1982
).
The effects of predation by wading birds (Ardeidae) and blue crabs (Callinectes sapidus) on the population size structure of the common mummichog, Fundulus heteroclitus
.
Estuar. Coast. Shelf Sci.
14
,
159
-
165
.
Kokurewicz
,
T.
(
2004
).
Sex and age related habitat selection and mass dynamics of Daubenton's bats Myotis daubentonii (Kuhl, 1817) hibernating in natural conditions
.
Acta Chiropt.
6
,
121
-
144
.
Kraemer
,
L. D.
and
Schulte
,
P. M.
(
2004
).
Prior PCB exposure suppresses hypoxia-induced up-regulation of glycolytic enzymes in Fundulus heteroclitus
.
Comp. Biochem. Physiol. C
139
,
23
-
29
.
Kramer
,
D. L.
(
1983
).
Aquatic surface respiration in the fishes of Panama: distribution in relation to risk of hypoxia
.
Env. Biol. Fish.
8
,
49
-
54
.
Kramer
,
D. L.
and
McClure
,
M.
(
1982
).
Aquatic surface respiration, a widespread adaptation to hypoxia in tropical freshwater fishes
.
Env. Biol. Fish.
7
,
47
-
55
.
Kramer
,
D. L.
and
Mehegan
,
J. P.
(
1981
).
Aquatic surface respiration, an adaptive response to hypoxia in the guppy, Poecilia reticulata (Pisces, Poeciliidae)
.
Env. Biol. Fish.
6
,
299
-
313
.
Lai
,
J. C. C.
,
Kakuta
,
I.
,
Mok
,
H. O. L.
,
Rummer
,
J. L.
and
Randall
,
D.
(
2006
).
Effects of moderate and substantial hypoxia on erythropoietin levels in rainbow trout kidney and spleen
.
J. Exp. Biol.
209
,
2734
-
2738
.
Lamont
,
P. A.
and
Gage
,
J. D.
(
2000
).
Morphological responses of macrobenthic polychaetes to low oxygen on the Oman continental slope, NW Arabian Sea
.
Deep Sea Res. II
47
,
9
-
24
.
Lanszki
,
J.
,
Molnár
,
M.
and
Molnár
,
T.
(
2006
).
Factors affecting the predation of otter (Lutra lutra) on European pond turtle (Emys orbicularis)
.
J. Zool.
270
,
219
-
226
.
Lau
,
G. Y.
,
Mandic
,
M.
and
Richards
,
J. G.
(
2017
).
Evolution of cytochrome c oxidase in hypoxia tolerant sculpins (Cottidae, Actinopterygii)
.
Mol. Biol. Evol.
34
,
2153
-
2162
.
Lefevre
,
S.
,
Stecyk
,
J. A. W.
,
Torp
,
M.-K.
,
Løvold
,
L. Y.
,
Sørensen
,
C.
,
Johansen
,
I. B.
,
Stensløkken
,
K.-O.
,
Couturier
,
C. S.
,
Sloman
,
K. A.
and
Nilsson
,
G. E.
(
2017
).
Re-oxygenation after anoxia induces brain cell death and memory loss in the anoxia-tolerant crucian carp
.
J. Exp. Biol.
220
,
3883
-
3895
.
LeMoine
,
C. M. R.
,
Bucking
,
C.
,
Craig
,
P. M.
and
Walsh
,
P. J.
(
2014
).
Divergent hypoxia tolerance in adult males and females of the plainfin midshipman (Porichthys notatus)
.
Physiol. Biochem. Zool.
87
,
325
-
333
.
Levin
,
L. A.
(
2003
).
Oxygen minimum zone benthos: adaptation and community response to hypoxia
.
Oceanog. Mar. Biol. Ann. Rev.
41
,
1
-
45
.
Lewis
,
J. M.
,
Costa
,
I.
,
Val
,
A. L.
,
Almeida-Val
,
V. M. F.
,
Gamperl
,
A. K.
and
Driedzic
,
W. R.
(
2007
).
Responses to hypoxia and recovery: repayment of oxygen debt is not associated with compensatory protein synthesis in the Amazonian cichlid, Astronotus ocellatus
.
J. Exp. Biol.
210
,
1935
-
1943
.
Losos
,
J. B.
(
2011
).
Convergence, adaptation, and constraint
.
Evolution
65
,
1827
-
1840
.
Love
,
J. W.
and
Rees
,
B. B.
(
2002
).
Seasonal differences in hypoxia tolerance in gulf killifish, Fundulus grandis (Fundulidae)
.
Environ. Biol. Fish.
63
,
103
-
115
.
Lushchak
,
V. I.
,
Bahnjukova
,
T. V.
and
Storey
,
K. B.
(
1998
).
Effect of hypoxia on the activity and binding of glycolytic and associated enzymes in sea scorpion tissues
.
Braz. J. Med. Biol. Res.
31
,
1059
-
1067
.
MacCormack
,
T. J.
,
Lewis
,
J. M.
,
Almeida-Val
,
V. M. F.
,
Val
,
A. L.
and
Driedzic
,
W. R.
(
2006
).
Carbohydrate management, anaerobic metabolism, and adenosine levels in the armoured catfish, Liposarcus pardalis (Castelnau), during hypoxia
.
J. Exp. Zool.
305A
,
363
-
375
.
Mandic
,
M.
,
Todgham
,
A. E.
and
Richards
,
J. G.
(
2009a
).
Mechanisms and evolution of hypoxia tolerance in fish
.
Proc. R. Soc. B Biol. Sci.
276
,
735
-
744
.
Mandic
,
M.
,
Sloman
,
K. A.
and
Richards
,
J. G.
(
2009b
).
Escaping to the surface: a phylogenetically independent analysis of hypoxia-induced respiratory behaviors in sculpins
.
Physiol. Biochem. Zool.
82
,
730
-
738
.
Mandic
,
M.
,
Speers-Roesch
,
B.
and
Richards
,
J. G.
(
2013
).
Hypoxia tolerance in sculpins is associated with high anaerobic enzyme activity in brain but not in liver or muscle
.
Physiol. Biochem. Zool.
86
,
92
-
105
.
Martin
,
K. L. M.
(
1991
).
Facultative aerial respiration in an intertidal sculpin, Clinocottus analis (Scorpaeniformes: Cottidae)
.
Physiol. Zool.
64
,
1341
-
1355
.
Martin
,
K. L. M.
(
1995
).
Time and tide wait for no fish: intertidal fishes out of water
.
Environ. Biol. Fish.
44
,
165
-
181
.
Martin
,
K. L. M.
(
1996
).
An ecological gradient in air-breathing ability among marine cottid fishes
.
Physiol. Zool.
69
,
1096
-
1113
.
Martinez
,
M. L.
,
Landry
,
C.
,
Boehm
,
R.
,
Manning
,
S.
,
Cheek
,
A. O.
and
Rees
,
B. B.
(
2006
).
Effects of long-term hypoxia on enzymes of carbohydrate metabolism in the Gulf killifish, Fundulus grandis
.
J. Exp. Biol.
209
,
3851
-
3861
.
Matey
,
V.
,
Richards
,
J. G.
,
Wang
,
Y.
,
Wood
,
C. M.
,
Rogers
,
J.
,
Davies
,
R.
,
Murray
,
B. W.
,
Chen
,
X.-Q.
,
Du
,
J.
and
Brauner
,
C. J.
(
2008
).
The effect of hypoxia on gill morphology and ionoregulatory status in the Lake Qinghai scaleless carp, Gymnocypris przewalskii
.
J. Exp. Biol.
211
,
1063
-
1074
.
Mathers
,
K. E.
,
Cox
,
J. A.
,
Wang
,
Y.
and
Moyes
,
C. D.
(
2014
).
Exploring the consequences of mitochondrial differences arising through hybridization of sunfish
.
Comp. Biochem. Physiol. A Mol. Integr. Physiol.
178
,
1
-
6
.
Mathias
,
J. A.
and
Barica
,
J.
(
1980
).
Factors controlling oxygen depletion in ice-covered lakes
.
Can. J. Fish. Aquat. Sci.
37
,
185
-
194
.
McBryan
,
T. L.
,
Healy
,
T. M.
,
Haakons
,
K. L.
and
Schulte
,
P. M.
(
2016
).
Warm acclimation improves hypoxia tolerance in Fundulus heteroclitus
.
J. Exp. Biol.
219
,
474
-
484
.
McKenzie
,
D. J.
,
Piraccini
,
G.
,
Steffensen
,
J. F.
,
Bolis
,
C. L.
,
Bronzi
,
P.
and
Taylor
,
E. W.
(
1995
).
Effects of diet on spontaneous locomotor activity and oxygen consumption in Adriatic sturgeon (Acipenser naccarii)
.
Fish Physiol. Biochem.
14
,
341
-
355
.
Mittelbach
,
G. G.
(
1984
).
Predation and resource partitioning in two sunfishes (Centrarchidae)
.
Ecology
65
,
499
-
513
.
Mulvey
,
J. M.
and
Renshaw
,
G. M. C.
(
2000
).
Neuronal oxidative hypometabolism in the brainstem of the epaulette shark (Hemiscyllium ocellatum) in response to hypoxic pre-conditioning
.
Neurosci. Lett.
290
,
1
-
4
.
Muusze
,
B.
,
Marcon
,
J.
,
van den Thillart
,
G.
and
Almeida-Val
,
V.
(
1998
).
Hypoxia tolerance of Amazon fish: respirometry and energy metabolism of the cichlid Astronotus oceallatus
.
Comp. Biochem. Physiol. A
120
,
151
-
156
.
Nasby-Lucas
,
N.
,
Dewar
,
H.
,
Lam
,
C. H.
,
Goldman
,
K. J.
and
Domeier
,
M. L.
(
2009
).
White shark offshore habitat: a behavioral and environmental characterization of the eastern Pacific shared offshore foraging area
.
PLoS ONE
4
,
e8163
.
Nelson
,
S. J.
(
2006
).
Fishes of the World
, 4th edn.
Hoboken, NJ
:
John Wiley & Sons
.
Nelson
,
J. A.
(
2016
).
Oxygen consumption rate v. rate of energy utilization of fishes: a comparison and brief history of the two measurements
.
J. Fish Biol.
88
,
10
-
25
.
Nilsson
,
G. E.
and
Ostlund-Nilsson
,
S.
(
2004
).
Hypoxia in paradise: widespread hypoxia tolerance in coral reef fishes
.
Proc. R. Soc. Lond. B
271
,
S30
-
S33
.
Nilsson
,
G. E.
and
Ostlund-Nilsson
,
S.
(
2006
).
Hypoxia tolerance in coral reef fishes
. In
Fish Physiology
, Vol.
21
(ed.
A. L.
Val
,
V. M. F.
Almeida-Val
and
D. J.
Randall
), pp.
583
-
596
.
London
:
Elsevier
.
Nilsson
,
G. E.
and
Östlund-Nilsson
,
S.
(
2008
).
Does size matter for hypoxia tolerance in fish?
Biol. Rev. Camb. Philos. Soc.
83
,
173
-
189
.
Nilsson
,
G. E.
and
Randall
,
D. J.
(
2010
).
Adaptations to hypoxia in fishes
. In
Respiratory Physiology of Vertebrates: Life With and Without Oxygen
(ed.
G.E.
Nilsson
), pp.
131
-
173
.
Cambridge
:
Cambridge University Press
.
Nilsson
,
G. E.
,
Rosen
,
P. R.
and
Johansson
,
D.
(
1993
).
Anoxic depression of spontaneous locomotor activity in crucian carp quantified by a computerized imaging technique
.
J. Exp. Biol.
180
,
153
-
162
.
Nilsson
,
G. E.
,
Hobbs
,
J.-P.
,
Munday
,
P. L.
and
Ostlund-Nilsson
,
S.
(
2004
).
Coward or braveheart: extreme habitat fidelity through hypoxia tolerance in a coral-dwelling goby
.
J. Exp. Biol.
207
,
33
-
39
.
Nilsson
,
G. E.
,
Hobbs
,
J.-P. A.
and
Östlund-Nilsson
,
S.
(
2007
).
Tribute to P.L. Lutz: respiratory ecophysiology of coral-reef teleosts
.
J. Exp. Biol.
210
,
1673
-
1686
.
Olofsson
,
M.
,
Vallin
,
A.
,
Jakobsson
,
S.
and
Wiklund
,
C.
(
2011
).
Winter predation on two species of hibernating butterflies: monitoring rodent attacks with infrared cameras
.
Anim. Behav.
81
,
529
-
534
.
Pedersen
,
C. L.
(
1987
).
Energy budgets for juvenile rainbow trout at various oxygen concentrations
.
Aquaculture
62
,
289
-
298
.
Perry
,
S. F.
,
Jonz
,
M. G.
and
Gilmour
,
K. M.
(
2009
).
Oxygen sensing and the hypoxic ventilatory response
. In
Fish Physiology. Volume 27: Hypoxia
(ed.
J. G.
Richards
,
A. P.
Farrell
and
C. J.
Brauner
), pp.
193
-
253
.
London
:
Elsevier
.
Popov
,
V. I.
,
Bocharova
,
L. S.
and
Bragin
,
A. G.
(
1992
).
Repeated changes of dendritic morphology in the hippocampus of ground squirrels in the course of hibernation
.
Neuroscience
48
,
45
-
51
.
Randle
,
A. M.
and
Chapman
,
L. J.
(
2004
).
Habitat use by the African anabantid fish Ctenopoma muriei: implications for costs of air breathing
.
Ecol. Freshw. Fish
13
,
37
-
45
.
Regan
,
M. D.
and
Richards
,
J. G.
(
2017
).
Rates of hypoxia induction alter mechanisms of O2 uptake and the critical O2 tension of goldfish
.
J. Exp. Biol.
220
,
2536
-
2544
.
Regan
,
M. D.
,
Gosline
,
J. M.
and
Richards
,
J. G.
(
2013
).
A simple and affordable calorespirometer for assessing the metabolic rates of fishes
.
J. Exp. Biol.
216
,
4507
-
4513
.
Regan
,
M. D.
,
Gill
,
I. S.
and
Richards
,
J. G.
(
2017a
).
Calorespirometry reveals that goldfish prioritize aerobic metabolism over metabolic rate depression in all but near-anoxic environments
.
J. Exp. Biol.
220
,
564
-
572
.
Regan
,
M. D.
,
Gill
,
I. S.
and
Richards
,
J. G.
(
2017b
).
Metabolic depression and the evolution of hypoxia tolerance in threespine stickleback, Gasterosteus aculeatus
.
Biol. Lett.
13
,
20170392
.
Renshaw
,
G. M. C.
,
Kerrisk
,
C. B.
and
Nilsson
,
G. E.
(
2002
).
The role of adenosine in the anoxic survival of the epaulette shark, Hemiscyllium ocellatum
.
Comp. Biochem. Physiol. B
131
,
133
-
141
.
Richards
,
J. G.
(
2009
).
Metabolic and molecular responses of fish to hypoxia
. In
Fish Physiology
, Vol.
27
(ed.
J. G.
Richards
,
A. P.
Farrell
and
C. J.
Brauner
), pp.
443
-
485
.
London
:
Elsevier
.
Richards
,
J. G.
(
2011
).
Physiological, behavioral and biochemical adaptations of intertidal fishes to hypoxia
.
J. Exp. Biol.
214
,
191
-
199
.
Richards
,
J. G.
,
Wang
,
Y. S.
,
Brauner
,
C. J.
,
Gonzalez
,
R. J.
,
Patrick
,
M. L.
,
Schulte
,
P. M.
,
Choppari-Gomes
,
A. R.
,
Almeida-Val
,
V. M.
and
Val
,
A. L.
(
2007
).
Metabolic and ionoregulatory responses of the Amazonian cichlid, Astronotus ocellatus, to severe hypoxia
.
J. Comp. Physiol. B
177
,
361
-
374
.
Richards
,
J. G.
,
Sardella
,
B. A.
and
Schulte
,
P. M.
(
2008
).
Regulation of pyruvate dehydrogenase in the common killifish, Fundulus heteroclitus, during hypoxia exposure
.
Am. J. Physiol. Regul. Integr. Comp. Physiol.
295
,
R979
-
R990
.
Roberts
,
J. J.
,
Höök
,
T. O.
,
Ludsin
,
S. A.
,
Pothoven
,
S. A.
,
Vanderploeg
,
H. A.
and
Brandt
,
S. B.
(
2009
).
Effects of hypolimnetic hypoxia on foraging and distributions of Lake Erie yellow perch
.
J. Exp. Mar. Biol. Ecol.
381
,
S132
-
S142
.
Rosa
,
R.
and
Seibel
,
B. A.
(
2010
).
Metabolic physiology of the Humboldt squid, Dosidicus gigas: implications for vertical migration in a pronounced oxygen minimum zone
.
Prog. Oceanogr.
86
,
72
-
80
.
Rosenblum
,
E. B.
,
Parent
,
C. E.
and
Brandt
,
E. E.
(
2014
).
The molecular basis of phenotypic convergence
.
Annu. Rev. Ecol. Evol. Syst.
45
,
203
-
226
.
Routley
,
M. H.
,
Nilsson
,
G. E.
and
Renshaw
,
G. M. C.
(
2002
).
Exposure to hypoxia primes the respiratory and metabolic responses of the epaulette shark to progressive hypoxia
.
Comp. Biochem. Physiol. A
131
,
313
-
321
.
Rowe
,
D. K.
and
Chisnall
,
B. L.
(
1995
).
Effects of oxygen, temperature and light gradients on the vertical distribution of rainbow trout, Oncorhynchus mykiss, in two North Island, New Zealand, lakes differing in trophic status
.
New Zeal. J. Mar. Fresh.
29
,
421
-
434
.
Saint-Paul
,
U.
(
1984
).
Physiological adaptation to hypoxia of a neotropical characoid fish Colossoma macropomum, Serrasalmidae
.
Environ. Biol. Fish.
11
,
53
-
62
.
Sanders
,
N. K.
and
Childress
,
J. J.
(
1990a
).
A Comparison of the respiratory function of the haemocyanins of vertically migrating and non-migrating pelagic, deep-sea oplophorid) shrimps
.
J. Exp. Biol.
152
,
167
-
187
.
Sanders
,
N. K.
and
Childress
,
J. J.
(
1990b
).
Adaptations to the deep-sea oxygen minimum layer: oxygen binding by the hemocyanin of the bathypelagic mysid, Gnathophausia ingens Dohrn
.
Biol. Bull.
178
,
286
-
294
.
Schmidtko
,
S.
,
Stramma
,
L.
and
Visbeck
,
M.
(
2017
).
Decline in global oceanic oxygen content during the past five decades
.
Nature
542
,
335
-
339
.
Schurmann
,
H.
and
Steffensen
,
J.
(
1994
).
Spontaneous swimming activity of atlantic cod gadus morhua exposed to graded hypoxia at three temperatures
.
J. Exp. Biol.
197
,
129
-
142
.
Scott
,
G. R.
,
Wood
,
C. M.
,
Sloman
,
K. A.
,
Iftikar
,
F. I.
,
De Boeck
,
G.
,
Almeida-Val
,
V. M. F.
and
Val
,
A. L.
(
2008
).
Respiratory responses to progressive hypoxia in the Amazonian Oscar, Astronotus ocellatus
.
Resp. Physiol. Neurobiol.
162
,
109
-
116
.
Seibel
,
B. A.
(
2011
).
Critical oxygen levels and metabolic suppression in oceanic oxygen minimum zones
.
J. Exp. Biol.
214
,
326
-
336
.
Seibel
,
B. A.
(
2013
).
The jumbo squid, Dosidicus gigas (Ommastrephidae), living in oxygen minimum zones II: blood–oxygen binding
.
Deep Sea Res. II Top. Stud. Oceanogr.
95
,
139
-
144
.
Seibel
,
B. A.
and
Drazen
,
J. C.
(
2007
).
The rate of metabolism in marine animals: environmental constraints, ecological demands and energetic opportunities
.
Philos. Trans. R. Soc. Lond. B
362
,
2061
-
2078
.
Seibel
,
B. A.
,
Thuesen
,
E. V.
and
Childress
,
J. J.
(
2000
).
Light-limitation on predator-prey interactions: consequences for metabolism and locomotion of deep-sea cephalopods
.
Biol. Bull.
198
,
284
-
298
.
Seibel
,
B. A.
,
Schneider
,
J. L.
,
Kaartvedt
,
S.
,
Wishner
,
K. F.
and
Daly
,
K. L.
(
2016
).
Hypoxia tolerance and metabolic suppression in oxygen minimum zone euphausiids: implications for ocean deoxygenation and biogeochemical cycles
.
Integr. Comp. Biol.
56
,
510
-
523
.
Shingles
,
A.
,
McKenzie
,
D. J.
,
Claireaux
,
G.
and
Domenici
,
P.
(
2005
).
Reflex cardioventilatory responses to hypoxia in the flathead gray mullet (Mugil cephalus) and their behavioral modulation by perceived threat of predation and water turbidity
.
Physiol. Biochem. Zool.
78
,
744
-
755
.
Shoubridge
,
E. A.
and
Hochachka
,
P. W.
(
1980
).
Ethanol: novel end product of vertebrate anaerobic metabolism
.
Science
209
,
308
-
309
.
Sloman
,
K. A.
,
Mandic
,
M.
,
Todgham
,
A. E.
,
Fangue
,
N. A.
,
Subrt
,
P.
and
Richards
,
J. G.
(
2008
).
The response of the tidepool sculpin, Oligocottus maculosus, to hypoxia in laboratory, mesocosm and field environments
.
Comp. Biochem. Physiol.
149
,
284
-
292
.
Sloman
,
K. A.
,
Sloman
,
R. D.
,
De Boeck
,
G.
,
Scott
,
G. R.
,
Iftikar
,
F. I.
,
Wood
,
C. M.
,
Almeida-Val
,
V. M. F.
and
Val
,
A. L.
(
2009
).
The role of size in synchronous air breathing of Hoplosternum littorale
.
Physiol. Biochem. Zool.
82
,
625
-
634
.
Smith
,
V. H.
,
Joye
,
S. B.
and
Howarth
,
R. W.
(
2006
).
Eutrophication of freshwater and marine ecosystems
.
Limnol. Oceanogr.
51
,
351
-
355
.
Soares
,
M. G. M.
,
Menezes
,
N. A.
and
Junk
,
W. J.
(
2006
).
Adaptations of fish species to oxygen depletion in a central Amazonian floodplain lake
.
Hydrobiologia
568
,
353
-
367
.
Sollid
,
J.
,
Angelis
,
P.
,
Gundersen
,
K.
and
Nilsson
,
G. E.
(
2003
).
Hypoxia induces adaptive and reversible gross morphological changes in crucian carp gills
.
J. Exp. Biol.
206
,
3667
-
3673
.
Sommer
,
R. S.
,
Niederle
,
M.
,
Labes
,
R.
and
Zoller
,
H.
(
2009
).
Bat predation by the barn owl Tyto alba in a hibernation site of bats
.
Folia Zool.
58
,
98
-
103
.
Speers-Roesch
,
B.
,
Richards
,
J. G.
,
Brauner
,
C. J.
,
Farrell
,
A. P.
,
Hickey
,
A. J. R.
,
Wang
,
Y. S.
and
Renshaw
,
G. M. C.
(
2012a
).
Hypoxia tolerance in elasmobranchs. I. Critical oxygen tension as a measure of blood oxygen transport during hypoxia exposure
.
J. Exp. Biol.
215
,
93
-
102
.
Speers-Roesch
,
B.
,
Brauner
,
C. J.
,
Farrell
,
A. P.
,
Hickey
,
A. J. R.
,
Renshaw
,
G. M. C.
,
Wang
,
Y. S.
and
Richards
,
J. G.
(
2012b
).
Hypoxia tolerance in elasmobranchs. II. Cardiovascular function and tissue metabolic responses during progressive and relative hypoxia exposures
.
J. Exp. Biol.
215
,
103
-
114
.
Speers-Roesch
,
B.
,
Mandic
,
M.
,
Groom
,
D. J. E.
and
Richards
,
J. G.
(
2013
).
Critical oxygen tensions as predictors of hypoxia tolerance and tissue metabolic responses during hypoxia exposure in fishes
.
J. Exp. Mar. Biol. Ecol.
449
,
239
-
249
.
Steffensen
,
J. F.
(
1985
).
The transition between branchial pumping and ram ventilation in fishes: energetic consequences and dependence on water oxygen tension
.
J. Exp. Biol.
114
,
141
-
150
.
Stenslokken
,
K.-O.
,
Sundin
,
L.
,
Renshaw
,
G. M. C.
and
Nilsson
,
G. E.
(
2004
).
Adenosinergic and cholinergic control mechanisms during hypoxia in the epaulette shark (Hemischyllium ocellatum), with emphasis on branchial circulation
.
J. Exp. Biol.
207
,
4451
-
4461
.
Stierhoff
,
K. L.
,
Targett
,
T. E.
and
Grecay
,
P. A.
(
2003
).
Hypoxia tolerance of the mummichog: the role of access to the water surface
.
J. Fish. Biol.
63
,
580
-
592
.
Sundin
,
L.
,
Nilsson
,
G. E.
,
Block
,
M.
and
Lofman
,
C. O.
(
1995
).
Control of gill filament blood-flow by serotonin in the rainbow trout, Oncorhynchus mykiss
.
Am. J. Physiol.
268
,
R1224
-
R1229
.
Tetens
,
V.
and
Lykkeboe
,
G.
(
1981
).
Blood respiratory properties of rainbow trout, Salmo gairdneri: responses to hypoxia acclimation and anoxic incubation of blood in vitro
.
J. Comp. Physiol.
145
,
117
-
125
.
Thomas
,
P.
,
Rahman
,
M. S.
,
Kummer
,
J. A.
and
Lawson
,
S.
(
2006
).
Reproductive endocrine dysfunction in Atlantic croaker exposed to hypoxia
.
Mar. Environ. Res.
62
,
249
-
252
.
Torres
,
J. J.
,
Aarset
,
A. V.
,
Donnelly
,
J.
,
Hopkins
,
T. L.
,
Lancraft
,
T. M.
and
Ainley
,
D. G.
(
1994
).
Metabolism of Antarctic micronektonic Crustacea as a function of depth of occurrence and season
.
Mar. Ecol. Prog. Ser.
113
,
207
-
219
.
Torres
,
J. J.
,
Grigsby
,
M. D.
and
Clarke
,
M. E.
(
2012
).
Aerobic and anaerobic metabolism in oxygen minimum layer fishes: the role of alcohol dehydrogenase
.
J. Exp. Biol.
215
,
1905
-
1914
.
Truchot
,
J.-P.
and
Duhamel-Jouve
,
A.
(
1980
).
Oxygen and carbon dioxide in the marine intertidal environment: diurnal and tidal changes in rockpools
.
Resp. Physiol.
39
,
241
-
254
.
Trueblood
,
L. A.
and
Seibel
,
B. A.
(
2013
).
The jumbo squid, Dosidicus gigas (Ommastrephidae), living in oxygen minimum zones I: oxygen consumption rates and critical oxygen partial pressures
.
Deep Sea Res. II Top. Stud. Oceanogr.
95
,
218
-
224
.
Turko
,
A. J.
,
Robertson
,
C. E.
,
Bianchini
,
K.
,
Freeman
,
M.
and
Wright
,
P. A.
(
2014
).
The amphibious fish Kryptolebias marmoratus uses different strategies to maintain oxygen delivery during aquatic hypoxia and air exposure
.
J. Exp. Biol.
217
,
3988
-
3995
.
Tzaneva
,
V.
,
Gilmour
,
K. M.
and
Perry
,
S. F.
(
2011
).
Respiratory responses to hypoxia or hypercapnia in goldfish (Carassius auratus) experiencing gill remodelling
.
Resp. Physiol. Neurobiol.
175
,
112
-
120
.
Ultsch
,
G. R.
(
1989
).
Ecology and physiology of hibernation and overwintering among freshwater fishes, turtles, and snakes
.
Biol. Rev.
64
,
435
-
515
.
Val
,
A. L.
and
Almeida-Val
,
V. M. F.
(
1995
).
Fishes of the Amazon and their Environments: Physiological and Biochemical Aspects
.
Berlin
,
Heidelberg
:
Springer-Verlag
.
Val
,
A. L.
,
Silva
,
M. N. P.
and
Almeida-Val
,
V. M. F.
(
1998
).
Hypoxia adaptation in fish of the Amazon: a never-ending task
.
S. Afr. J. Zool.
33
,
107
-
114
.
van Ginneken
,
V.
and
van den Thillart
,
G.
(
2009
).
Metabolic depression in fish measured by direct calorimetry: a review
.
Thermochimica Acta
483
,
1
-
7
.
van Raaij
,
M. T. M.
,
van den Thillart
,
G. E. E. J. M.
,
Vianen
,
G. J.
and
Steffens
,
A. B.
(
1996
).
Substrate mobilization and hormonal changes in rainbow trout (Oncorhynchus mykiss, L.) and common carp (Cyprinus carpio, L.) during deep hypoxia and subsequent recovery
.
J. Comp. Physiol.
166
,
443
-
452
.
Vetter
,
R.
,
Kohin
,
S.
,
Preti
,
A.
and
McClatchie
,
S.
(
2008
).
Predatory interactions and niche overlap between mako shark, Isurus oxyrinchus, and jumbo squid, Dosidicus gigas, in the California Current
.
Calif. Coop. Ocean Fish Invest. Rep.
49
,
142
-
156
.
Virani
,
N. A.
and
Rees
,
B. B.
(
2000
).
Oxygen consumption, blood lactate and inter-individual variation in the gulf killifish, Fundulus grandis, during hypoxia and recovery
.
Comp. Biochem. Physiol. A
126
,
397
-
405
.
Vornanen
,
M.
(
2004
).
Seasonality of dihydropyridine receptor binding in the heart of an anoxia-tolerant vertebrate, the crucian carp (Carassius carassius L.)
.
Am. J. Physiol.
287
,
R1263
-
R1269
.
Vornanen
,
M.
,
Stecyk
,
J. A. W.
and
Nilsson
,
G. E.
(
2009
).
The anoxia-tolerant crucian carp (Carassius carassius L.)
. In
Fish Physiology
Vol.
27
(ed.
J. G.
Richards
,
A. P.
Farrell
and
C. J.
Brauner
), pp.
397
-
441
. London:
Elsevier
.
Vulesevic
,
B.
and
Perry
,
S. F.
(
2006
).
Developmental plasticity of ventilatory control in zebrafish, Danio rerio
.
Resp. Physiol. Neurobiol
154
,
396
-
405
.
Wang
,
T.
,
Lefevre
,
S.
,
Thanh Huong
,
D. T.
,
Cong
,
N. V.
and
Bayley
,
M.
(
2009
).
The effects of hypoxia on growth and digestion
. In
Fish Physiology
, Vol.
27
(ed.
J. G.
Richards
,
A. P.
Farrell
and
C. J.
Brauner
), pp.
361
-
396
.
London
:
Elsevier
.
Wannamaker
,
C. M.
and
Rice
,
J. A.
(
2000
).
Effects of hypoxia on movements and behavior of selected estuarine organisms from the southeastern United States
.
J. Exp. Mar. Biol. Ecol.
249
,
145
-
163
.
Weibel
,
R. R.
(
1984
).
The Pathway for Oxygen: Structure and Function in the Mammalian Respiratory System
.
Cambridge, MA
:
Harvard University Press
.
Wiklund
,
C.
,
Vallin
,
A.
,
Friberg
,
M.
and
Jakobsson
,
S.
(
2008
).
Rodent predation on hibernating peacock and small tortoiseshell butterflies
.
Behav. Evol. Sociobiol.
62
,
379
-
389
.
Wise
,
G.
,
Mulvey
,
J. M.
and
Renshaw
,
G. M. C.
(
1998
).
Hypoxia tolerance in the epaulette shark (Hemiscyllium ocellatum)
.
J. Exp. Zool.
281
,
1
-
5
.
Wong
,
C. C.
,
Drazen
,
J. C.
,
Callan
,
C. K.
and
Korsmeyer
,
K. E.
(
2018
).
Hypoxia tolerance in coral-reef triggerfishes (Balistidae)
.
Coral Reefs
37
,
215
-
225
.
Wright
,
W. G.
and
Raymond
,
J. A.
(
1978
).
Air-breathing in a California sculpin
.
J. Exp. Zool.
203
,
171
-
176
.
Wu
,
R. S. S.
(
2009
).
Effects of hypoxia on fish reproduction and development
. In
Fish Physiology
, Vol.
27
(ed.
J. G.
Richards
,
A. P.
Farrell
and
C.J.
Brauner
), pp.
79
-
141
.
London
:
Elsevier
.
Yang
,
T. H.
and
Somero
,
G. N.
(
1993
).
Effects of feeding and food deprivation on oxygen consumption, muscle protein concentration and activities of energy metabolism enzymes in muscle and brain of low-living (Scorpaena guttata) and deep-living (Sebastolobus alascanus) scorpaenid fishes
.
J. Exp. Biol.
181
,
213
-
232
.
Yoshiyama
,
R. M.
and
Cech
,
J. J.
Jr
. (
1994
).
Aerial respiration by rocky intertidal fishes of California and Oregon
.
Copeia
1994
,
153
-
158
.
Yoshiyama
,
R. M.
,
Valpey
,
C. J.
,
Schalk
,
L. L.
,
Oswald
,
N. M.
,
Vaness
,
K. K.
,
Lauritzen
,
D.
and
Limm
,
M.
(
1995
).
Differential propensities for aerial emergence in intertidal sculpins (Teleostei; Cottidae)
.
J. Exp. Mar. Biol. Ecol.
191
,
195
-
207
.

Competing interests

The authors declare no competing or financial interests.

Supplementary information